You are on page 1of 29

Inferring Causality from Noninvasive Brain Stimulation in

Cognitive Neuroscience

Til Ole Bergmann1 and Gesa Hartwigsen2

Abstract
■ Noninvasive brain stimulation (NIBS) techniques, such as and in connected remote sites of the network, which in conse-
transcranial magnetic stimulation or transcranial direct and al- quence affects the cognitive function of interest and eventually
ternating current stimulation, are advocated as measures to en- results in a change of the behavioral measure. Importantly, ev-
able causal inference in cognitive neuroscience experiments. ery link in this causal chain of effects can be confounded by
Transcending the limitations of purely correlative neuroimaging several factors that have to be experimentally eliminated or con-
measures and experimental sensory stimulation, they allow to trolled to attribute the observed results to their assumed cause.
experimentally manipulate brain activity and study its conse- This is complicated by the fact that many of the mediating and
quences for perception, cognition, and eventually, behavior. confounding variables are not directly observable and dose–
Although this is true in principle, particular caution is advised response relationships are often nonlinear. We will walk the
when interpreting brain stimulation experiments in a causal reader through the chain of causation for a generic cognitive
manner. Research hypotheses are often oversimplified, disre- neuroscience NIBS study, discuss possible confounds, and ad-
garding the underlying (implicitly assumed) complex chain of vise appropriate control conditions. If crucial assumptions are
causation, namely, that the stimulation technique has to gener- explicitly tested (where possible) and confounds are experi-
ate an electric field in the brain tissue, which then evokes or mentally well controlled, NIBS can indeed reveal cause–effect
modulates neuronal activity both locally in the target region relationships in cognitive neuroscience studies. ■

INTRODUCTION
were discovered as methods for subthreshold modula-
Noninvasive brain stimulation (NIBS) techniques, such as tion of neuronal activity and thus of cognitive function
TMS or transcranial direct and alternating current stimu- ( Vosskuhl, Strüber, & Herrmann, 2018; Yavari, Jamil,
lation (TDCS/TACS), allow to experimentally manipulate Mosayebi Samani, Vidor, & Nitsche, 2018). Although cog-
neuronal activity in the healthy human brain in a tempo- nitive neuroscience studies using these NIBS techniques
rally and spatially specific manner, thereby overcoming often claim to test the “causal relevance” of a specific
the merely correlative nature of electrophysiological brain region or neuronal activity pattern for a specific
and neuroimaging techniques (Bergmann, Karabanov, cognitive function or behavior, the underlying cause–
Hartwigsen, Thielscher, & Siebner, 2016). Their ability effect relationships are rarely made explicit. However,
to bypass sensory input channels and directly affect brain to justify such causal inference, the theoretically assumed
activity makes them unparalleled tools for studying chain of causation, leading from the applied stimulation
cause–effect relationships between neuronal activity to the observed behavioral change, has to hold for a con-
and cognitive function. Shortly after its invention crete empirical experiment. Crucially, every single cause–
(Barker, Jalinous, & Freeston, 1985), TMS was already effect link in this causal chain can be interrupted or
demonstrated to be capable of suppressing visual percep- confounded by several factors, which are best eliminated
tion (Amassian et al., 1989), and by now, the “disruptive” or controlled experimentally to attribute the observed
or “interfering” effects of TMS have a long-standing tradi- results to their assumed cause. We will start by briefly
tion in cognitive neuroscience, following the so-called outlining the core element of this article: a simplified
“virtual lesion” approach (Pascual-Leone, Walsh, & five-step chain of causation for cognitive neuroscience
Rothwell, 2000; Walsh & Cowey, 2000). Later, TDCS NIBS studies and its principal confounders. We will then
(Nitsche & Paulus, 2000) and TACS (Antal et al., 2008) introduce general experimental approaches using NIBS
and discuss the concept of causal inference for the case
of experimental NIBS studies in cognitive neuroscience
1
Leibniz Institute for Resilience Research, Mainz, Germany, before we walk the reader step-by-step through the
2 five-step chain of causation. Afterward, we will discuss
Max Planck Institute for Human Cognitive and Brain Sciences,
Leipzig, Germany potential confounders in more detail and review the

© 2020 by the Massachusetts Institute of Technology. Published under Journal of Cognitive Neuroscience X:Y, pp. 1–29
a Creative Commons Attribution 4.0 International (CC BY 4.0) license. https://doi.org/10.1162/jocn_a_01591
available experimental control conditions to counteract Figure 1B, we provide a more elaborated causal diagram,
them before we conclude by providing 12 general recom- but for the sake of comprehensibility, it is still simplified
mendations for designing valid NIBS studies. and reduced to the key variables (some of them summa-
Please note that, in the context of this article, causal in- rizing multiple smaller ones). The red elements connected
ference simply means “inferring causality” or “inferring by red arrows indicate the core chain of causation leading
that one variable is the cause of another” (Scheines, from application of NIBS to an observable behavioral ef-
2005), an inference that may either be based on the con- fect. The yellow box and arrows indicate the causal route
trolled randomized experiment or, under certain condi- by which task demands and the current brain state drive
tions, on observational data alone, when using the causal local and network activity and thereby the respective cog-
inference framework developed by Judea Pearl (Pearl & nitive function and behavioral responses. This causal path
Mackenzie, 2018; Pearl, 2010) and others. A more general is not only a source of confounding but is relevant for task
introduction into the latter and its implications for neuro- performance in the absence of NIBS and the main drive for
imaging studies is outside the scope of this article. While the cognitive function of interest. In fact, NIBS-induced
adopting elements from this particular causal inference brain activity per se is insufficient to cause more complex
framework, this review remains largely focused on the clas- cognitive functions and only modulates ongoing task-
sical experimentalist’s framework of inferring causation via related neuronal activity. The black arrows indicate addi-
experimental manipulation. We primarily aim to raise tional causal relationships that result from or eventually af-
awareness for the underlying (often implicitly assumed) fect elements of the main causal chain of interest and may
chain of causation in NIBS studies, their potential con- thereby produce confounding via central (blue boxes) and
founds, and respective experimental control measures, peripheral (green boxes) off-target stimulation. We segre-
encouraging the conduction of well-planned and well- gated the core chain of causation from NIBS application to
controlled NIBS experiments that actually justify causal behavior into five cause–effect pairs as described below.
inference, that is, the conclusion of cause–effect relation- Arrow 1: The application of NIBS produces an electric
ships between neuronal activity and cognitive function. field (E-field) in the brain tissue, either via transcranial
electro-magneto-electric induction (TMS) or the direct
administration of weak transcranial currents (transcranial
The Chain of Causation in NIBS Studies
current stimulation [TCS]).
A chain of causation (or causal chain) refers to an unin- Arrow 2: The E-field then interacts with the neurons’
terrupted concatenation of cause–effect pairs, leading membrane potential to immediately (online) evoke neu-
from an initial cause of interest via a number of mediating ronal firing (TMS) or modulate the threshold for sponta-
variables to an eventual effect of interest. Given that most neous firing (TCS) locally in the targeted brain region,
effects have many causes and themselves cause many ef- activating specific intracortical circuit motifs and asso-
fects, such a chain represents only one specific path ciated neuronal signatures ( Womelsdorf, Valiante, Sahin,
through an entire causal diagram. A causal diagram can Miller, & Tiesinga, 2014). In the case of offline TMS/TCS
be formalized as variables connected with arrows that in- protocols, it additionally triggers processes of synaptic
dicate causation (A ➔ B) instead of mere association (A, plasticity.
B), with the left and right variables representing cause Arrow 3: If driving output neurons of the targeted net-
and effect for a particular cause–effect pair (Pearl & work node to suprathreshold levels, local neuronal activ-
Mackenzie, 2018). Importantly, there is no ultimate cause ity transsynaptically spreads to other connected brain
or effect, and the partial chain to be considered depends regions of the targeted network via intercortical axonal
entirely on the research question at hand. Once the hy- projections, activating large-scale and remote circuit mo-
pothesis has defined cause and effect of interest, any in- tifs as well as remote changes in synaptic strength.
termediate element within a causal path connecting them Arrow 4: The immediate (online) or subsequent (off-
is referred to as “mediator.” In contrast, elements that are line) effects on local and large-scale circuit motifs can dis-
associated with the cause of interest (i.e., causing it or turb or facilitate the specific task-relevant neuronal
merely covarying with it but not caused by it) and that computations mediating the cognitive function of inter-
influence the effect of interest are considered con- est, which is reflected in either respective changes of
founders, because their influence on the effect of interest the outcome or the completion time of these processes.
is mixed with that of the cause of interest, preventing the Arrow 5: With regard to the motor responses exerted
straightforward attribution of causal influence to the lat- in the context of a specific behavioral task, these altered
ter and therefore the identification of an unambiguous cognitive processes eventually result in changed error
causal path through the causal diagram. rates or RTs.
The somewhat naive level at which hypotheses are often
phrased in NIBS studies is depicted in Figure 1A. The stim-
Confounding in NIBS Studies
ulation is expected to affect a single circumscribed brain
region, which has an effect on behavior, because the brain Most cognitive neuroscience studies investigate the neuro-
region is causally relevant for producing that behavior. In nal implementation of a cognitive function and therefore

2 Journal of Cognitive Neuroscience Volume X, Number Y


Figure 1. Causal diagram for NIBS studies in cognitive neuroscience. (A) Naive chain of causation stated for many NIBS studies: Stimulation is
expected to affect a single brain region, which has an effect on behavior, because the brain region is causally relevant for producing that behavior. (B)
A more elaborated causal diagram, with red arrows indicating the core chain of causation; yellow arrows, the impact of task demands and brain state;
and black arrows, additional causal relationships that may produce confounding. Application of NIBS produces an E-field in the brain tissue (1), which
evokes or modulates local neuronal activity in the target region (2), which then spreads via synaptic connections to other brain regions within the
same target network (3) and affects the task-relevant cognitive processes of interest mediated by the local target region or the target network (4),
which eventually results in a motor response as part of this task (5). However, NIBS also produces E-fields and thereby neuronal activity in nontarget
regions, spreading in nontarget networks and affecting cognitive processes of interest and of no interest (blue boxes), and thereby behavioral
outcome. NIBS also creates an E-field in the periphery, causing afferent input and neuronal activity in sensory regions (green boxes), affecting both
target and nontarget networks as well as cognitive processing. Importantly, task demands and the current brain state are the main drive of local and
network neuronal activity and thus engage the cognitive function itself, whereas NIBS-related brain activity is merely modulating task-related
neuronal activity and cognition.

hypothesize that neuronal activity in a certain brain region is relationships. Depending on the expertise and educational
causal to that cognitive function of interest. NIBS is merely background of the researcher, some of these links are typ-
considered a means to manipulate the cause “neuronal ically less thoroughly elaborated than others (e.g., the kind
activity” (via an E-field), whereas behavioral measures are of neuronal activity induced by the E-field or the precise
used as an observable proxy to assess the hidden effect behavioral changes to be expected from changes in cogni-
“cognitive function.” Every single link in this causal path tive function), crucial assumptions about mediators re-
can be confounded by several (known or unknown) factors main untested (e.g., whether the TMS-induced E-field
(yellow, blue, and green boxes in Figure 1B), which are actually produced the neuronal activity that was aimed
best experimentally eliminated or controlled for to rule for), and potential confounds are uncontrolled (e.g., the
out alternative explanations for the observed data and to sensory input because of peripheral costimulation).
draw strong conclusions regarding the hypothesized
cause–effect relationship. This is complicated by the facts
GENERAL APPROACHES FOR NIBS IN
that (i) many of the mediating and confounding variables
COGNITIVE NEUROSCIENCE
are not directly observable and have to be approximated
by imulation (E-field), neuroimaging (neuronal activity), Although NIBS is often depicted as a means by which we
or modeling (cognitive function) and (ii) few of the rele- can simply “switch off” or “knock out” entire brain regions
vant cause–effect pairs express linear dose–response or realistically mimic endogenous oscillatory activity, the

Bergmann and Hartwigsen 3


reality is more complicated, and the mechanism of action minutes to hours (Figure 2; for a more detailed discus-
and applicability of a specific technique need to be consid- sion of NIBS approaches, see Bergmann et al., 2016).
ered before choosing it to manipulate neuronal activity in a Online approaches, assessing the immediate neural re-
specific study. Although often lumped together as “NIBS,” sponse to stimulation, can be used to (i) quantify prop-
TMS and TCS rely on different neurophysiological princi- erties such as cortical excitability or connectivity and their
ples of action. Sharing the principal mediating mechanism modulation by brain state (including task engagement;
of an E-field being imposed on the brain tissue, the result- Figure 2A), (ii) interfere with ongoing task-related or
ing neuronal effects differ markedly. The fast-changing, spontaneous neuronal activity and thereby with cogni-
high-amplitude E-field gradients (< 300 μs) caused by tion (Figure 2B), or (iii) modulate, more gently, the level
TMS are sufficient to fully depolarize the membrane poten- (“gating”) and timing (“entrainment”) of neuronal activity
tial of cortical neurons, causing the immediate emergence and thereby cognitive function (Figure 2C). In contrast,
of action potentials (APs; suprathreshold stimulation). In offline approaches can be utilized to either (iv) inhibit
contrast, the much weaker constant or alternating E-fields or (v) facilitate neuronal excitability for an extended pe-
caused by TCS are assumed to merely shift the neurons’ riod via mechanisms of synaptic plasticity, assessing its
membrane potential slightly toward depolarization or hy- subsequent effects on neuronal activity and cognition
perpolarization, thus modulating the likelihood of APs to (Figure 2D and E). Although relying on different neuro-
emerge spontaneously (subthreshold stimulation). This physiological mechanisms (which we will discuss in a
fundamental difference has important consequences for later section), both online interference and offline inhibi-
the kind of experimental approaches suitable for TMS tion approaches have been referred to as “virtual lesion”
and TCS, respectively. Transcranial ultrasound stimulation and are most frequently used in cognitive neuroscience
is yet another promising NIBS technique for neuromodu- to test whether and when a cortical region is “causally
lation, which has recently received growing attention be- relevant” for a cognitive function. Online modulation
cause of its capability of stimulating very circumscribed additionally allows to investigate which neuronal patterns,
volumes deep in the brain while sparing the overlying for example, oscillatory frequencies, are mediating or
tissue, and can be expected to amend the NIBS toolbox supporting a specific cognitive function. Whereas TMS
for human applications in the near future (Fomenko, works for all these approaches, the much weaker TCS
Neudorfer, Dallapiazza, Kalia, & Lozano, 2018). Transcranial cannot be used for quantification (as it does not trigger
ultrasound stimulation is not based on the induction of APs) and hardly for interference. In contrast, it is optimally
E-fields but presumably involves the mechanic impact of suited to modulate without disruption and can produce
focused sound pressure waves on the neurons’ membrane lasting offline effects. We will see that net increases and
and/or ion channels ( Jerusalem et al., 2019) and will not decreases in cortical excitability (online as well as offline)
be discussed within the scope of this article. do not simply translate into respective behavioral improve-
ments and impairments.

Experimental Approaches Using NIBS


CAUSAL INFERENCE FROM NEUROIMAGING
Depending on their specific stimulation parameters, both
VS. NIBS IN COGNITIVE NEUROSCIENCE
TMS and TCS are able to exert not only immediate
(online) effects during stimulation but also subsequent Causal inference refers to the process of inferring cause–
(offline) effects that outlast the stimulation itself for effect relationships based on the observed changes of an

Figure 2. Schematic representation of noninvasive brain stimulation approaches. (A) Online quantification: a stimulation strong enough to cause a
direct output of the targeted region/network (with TMS, not TCS) that allows to quantify cortical excitability via MEPs or phosphene reports. (B)
Online interference: a disruption of ongoing task-related or spontaneous brain activity (with TMS, rather not TCS) that disturbs a cognitive function.
(C) Online modulation: a moderate modulation of the level (“gating” via low-intensity TMS or TDCS) or timing (“entrainment” via TACS or rhythmic
TMS) of neuronal activity that interacts with ongoing task-related or spontaneous neuronal activity without disrupting it. (D) Offline facilitation: an
increase in cortical excitability (triggered by rTMS or prolonged TCS) presumably mediated via LTP of the stimulated synapses. (E) Offline inhibition:
a decrease in cortical excitability (triggered by rTMS or prolonged TCS) presumably mediated via LTD of the stimulated synapses.

4 Journal of Cognitive Neuroscience Volume X, Number Y


effect, after changes of its hypothesized cause. In contrast nor directly be observed. We will thus use the terms
to inferring a mere association of the two variables by “cause” and “effect” within the causal inference frame-
means of conditional probabilities, that is, P(effect| work, but “IV” and “DV” when taking the experimentalist’s
cause), causal inference assigns a direction to their rela- perspective.
tionship, assuming that active manipulation of the cause
(experimentally or counterfactually), with everything else
Mapping Correlational Relationships with
held constant (“ceteris paribus”), produces the effect, but
Noninvasive Neuroimaging
not vice versa. This asymmetric relationship has been for-
malized via the do operator, where P(effect|cause) = P In neuroimaging studies, the cognitive function of inter-
(effect|do(cause)), whereas P (cause|effect) ≠ P(cause| est is isolated experimentally as the only difference be-
do(effect)) = P (cause) (Pearl & Mackenzie, 2018). A sim- tween respective task conditions (levels of the IV ),
ple example in the context of NIBS would be that TMS whereas the associated neural activity is assessed (as
(with sufficient intensity) of the primary motor cortex DV) and contrasted between conditions. By experimental
(M1) hand area causes a contralateral finger movement variation of the IV in randomized controlled trials, clever
∼20 msec later or at least increases its likelihood (P(con- design of task conditions, and careful experimental con-
traction|do(TMS)), but a spontaneous finger movement trol of possible confounders, such experiments, unlike
does not affect the likelihood of TMS to occur (P[TMS| mere observational studies, allow to attribute the ob-
do(contraction)] = P(TMS)). In fact, this asymmetry served DV change to variations of the IV. Nonetheless,
holds for every single cause–effect pair in the causal the causal direction between cognitive function and brain
chain mediating the effect from TMS pulse to finger activity is not easily derived from such experiments.
movement (i.e., TMS pulse ➔ E-field ➔ APs in cortical Commonly, this allocation of IV and DV does not imply
neurons in M1 Layer 2/3 or premotor cortex ➔ APs in cor- that a certain mental state is expected to cause the re-
ticospinal output neurons in M1 Layer 5 ➔ APs in spinal spective brain state (P[“brain state”|do(“mental state”)])
motoneurons ➔ muscular APs ➔ muscle contraction ➔ but rather the opposite (P[“mental state”|do(“brain
finger movement). For behavioral task performance in state”)]) or, at least, that the noncausal relation of super-
cognitive neuroscience studies, the causal chain is typi- venience is assumed (Dijkstra & de Bruin, 2016). While
cally even more complex, and for the causal diagram in trying to avoid taking any particular philosophical posi-
Figure 1B, the numerous single steps have been consol- tion with respect to the mind–body problem (for an in-
idated into a few categories for the sake of simplification. troduction, see Chambliss, 2018; Nagel, 1993), we will,
It has been argued that causal relationships can provi- for the ease of argument, assume here that a cognitive
sionally be inferred from observational data alone within function causally depends on a specific neuronal sub-
the causal inference framework when using the do- strate (i.e., the structure) and temporospatial patterns
calculus to decide which confounding variables should of neuronal activity it produces (P[“cognitive function”|
be statistically adjusted for and which should rather not do(“brain activity”)]). This should not be misunderstood
to avoid the introduction of spurious effects (Pearl & as a dualist view on mental causation, because “cognitive
Mackenzie, 2018; Pearl, 2010). This approach has gained function” does not refer to a conscious, phenomenal ex-
increasing interest in the functional neuroimaging com- perience or mental state but rather pragmatically to the
munity as well, where its applicability is fiercely debated mechanisms of information processing and the computa-
(Reid et al., 2019; Mehler & Kording, 2018; Grosse- tions that eventually give rise to a certain behavior, in the
Wentrup, Janzing, Siegel, & Schölkopf, 2016; Weichwald following generously spanning anything from perception,
et al., 2015; Ramsey et al., 2010). In contrast, the classical via higher-order cognitive processes, to motor function.
experiment solves this caveat elegantly also with respect From that perspective, a cognitive neuroimaging experi-
to unknown confounders via the randomized allocation ment thus engages the participant in a task requiring for
of levels of the independent variable (IV) to observational its completion the recruitment of certain neuronal net-
units across multiple experimental repetitions (i.e., do works and mechanisms implementing the respective tar-
[cause]), while observing the resulting changes in the de- get cognitive function, while measuring brain activity as a
pendent variable (DV) as a function of the IV level (i.e., function of task condition (for a detailed discussion, see
P(effect|do(cause)). Although manipulation of the IV Dijkstra & de Bruin, 2016). However, because not all as-
and resulting changes of the DV relate to cause and effect sociated brain activities may be causally contributing to
for the case of a randomized controlled trial with a single the cognitive function engaged by the task, the measured
cause–effect pair, it is more ambiguous for more complex neuronal activity cannot qualify unambiguously as its
chains of causation with multiple cause–effect pairs lined cause but merely as its neuronal correlate.
up (like in Figure 1B), where only the first cause (here,
application of stimulation) relates to the experimental
Mapping Causal Relationships with NIBS
manipulation of an IV and only the last effect (here, be-
havior) is assessed as DV, whereas all intermediate steps This ambiguity of causal direction can be resolved when
are hidden variables that can neither be manipulated using NIBS to experimentally manipulate brain activity

Bergmann and Hartwigsen 5


independent of task engagement instead of merely ob- step is far from trivial, and simply holding a TMS coil or
serving it via neuroimaging, while measuring behavioral attaching a TCS electrode over the assumed target region
task performance as a proxy for the integrity of a cogni- is not sufficient for many reasons.
tive function (Sack, 2006; for a detailed discussion, see
Dijkstra & de Bruin, 2016). Therefore, NIBS-related changes
in behavior can principally be interpreted as causal effects
Identifying the Target
of the experimentally induced change in brain activity
(P[“cognitive function”|do(“brain activity”)]). There are Before we can attempt to stimulate a specific target site of
few examples when NIBS alone is sufficient to produce interest, we need to determine its location. Depending on
motor behavior or a perceptual phenomenon, such as the the spatial specificity of the NIBS method, these targets are
well-established induction of muscle responses as quanti- easily underspecified or overspecified. For TMS, effective
fied by the motor-evoked potentials (MEPs) or phosphenes current densities are restricted to less than a cubic centi-
(illusory perceptions of light) after TMS of the primary meter (Brasil-Neto, McShane, Fuhr, Hallett, & Cohen,
motor and visual cortex, respectively (Kammer, 1998; 1992), and targets such as the “posterior parietal cortex”
Mills, Boniface, & Schubert, 1992). For these cases, the or “dorsolateral pFC” are very unspecific, given that (i)
online quantification approach (Figure 2A) can be em- the functional organization of most brain areas is topo-
ployed to demonstrate a clear causal relationship between graphically more fine-grained and (ii) only a portion of that
neuronal activity in the respective brain structure (e.g., anatomical structure will receive effective stimulation. For
TMS-induced firing of corticospinal motor neurons in M1) TCS, in contrast, even entire brain regions can hardly be
and behavioral outcome (e.g., contraction of a contralateral stimulated in isolation. Irrespective of the spatial specific-
hand muscle), and even dose–response relationships can ity of the NIBS technique, the target site can principally be
be identified (e.g., the sigmoidal function relating increas- determined based on (i) its function, (ii) its neuroanatom-
ing TMS intensity to increasing MEP amplitude). However, ical location, or (iii) even its location relative to the skull
with NIBS in humans, most other brain targets do not alone. A functional TMS localizer can be used to determine
result in overt outputs. Instead, behavioral tasks are re- a motor or phosphene hot spot based on MEPs and phos-
quired to engage the cognitive function of interest and phenes, respectively. This approach is highly specific as it
its neuronal correlate, while using NIBS to manipulate allows fine-tuned coil positioning based on the immediate
the activity or excitability of a target brain region before feedback from the output variables, ensuring that the
(offline; Figure 2D and E) or during (online; Figure 2B intended neuron population is effectively stimulated.
and C) task performance to reveal its causal contribution Unfortunately, this method is only available for very few
to this cognitive function. This brings us back to the heart targets (i.e., motor and visual cortex; see Bergmann
of this article: the chain of causation that is tacitly hypoth- et al., 2016). The second-best option is a localizer via
esized for NIBS studies in cognitive neuroscience and its fMRI, for example, to determine the FEF from a covert spa-
many possible confounders that complicate the causal tial attention task (Marshall, O’Shea, Jensen, & Bergmann,
interpretation of NIBS results (Figure 1B). In the next 2015) or the extrastriate body area from contrasting body
sections, we will walk the reader step-by-step through parts versus other objects (Zimmermann, Verhagen, de
the causal path and discuss for each step under which Lange, & Toni, 2016). While providing no information on
conditions a cause–effect relationship can be assumed. coil orientation, the target voxel can be determined based
Afterward, we will describe how these causal links can on individual statistical maps. For functions tightly linked
be confounded by variables that systematically covary to an identifiable anatomical location (e.g., motor hand
with the cause and constitute an alternative cause for its knob), an individual structural MRI scan alone may be used
effect. Only if an uninterrupted chain of causation can be to identify the target coordinates, yet allowing consider-
established without confounding causes for the individual able uncertainty within that area. When ignoring interindi-
links, the conclusion can be drawn that “neuronal activity vidual variability in a structure–function relationship,
in region X is causing behavior Y.” Montreal Neurological Institute coordinates from the liter-
ature can be utilized after transforming them to native
space with the help of an individual structural MRI
(Duecker et al., 2014). When disregarding individual brain
FROM NIBS APPLICATION TO E-FIELDS
anatomy altogether, the 10–20 EEG electrode system can
(ARROW 1)
be used to roughly estimate the location of specific brain
The first cause–effect pair (Arrow 1) is often implicitly as- regions (e.g., F3 for the left dorsolateral pFC or P4 for the
sumed without further discussion, namely, that the ap- right posterior parietal cortex). Systematic comparisons
plied NIBS technique produces an E-field of desired revealed that, with decreasing individualization across
intensity, extent, and direction in the target brain region, the above-described methods, the number of participants
without affecting nontarget brain regions and without the required to observe a significant effect increases dramati-
target brain region being inadvertently affected by other cally (Sack et al., 2009; Sparing, Buelte, Meister, Pauš, &
factors associated with the stimulation. This first crucial Fink, 2008). Importantly, all approaches besides the TMS

6 Journal of Cognitive Neuroscience Volume X, Number Y


localizer and the 10–20 system require an MR-informed FROM E-FIELDS TO LOCAL NEURONAL
frameless stereotactic neuronavigation setup to position EFFECTS (ARROW 2)
the TMS coil over the target site, which is considered
The second cause–effect pair (Arrow 2) refers to the im-
state-of-the-art to maintain coil position within and across
pact of the E-field on local neuronal activity. For the sake
experimental sessions.
of simplicity and to prevent overloading the causal dia-
gram in Figure 1B, local neuronal activity here also refers
to effects secondary to the initial neuronal response,
Reaching the Target such as the activation of local circuit motifs, shifts in neu-
ronal excitability, entrainment of local neuronal oscilla-
Once we know where to stimulate, how can we ensure
tors, and local synaptic plasticity. The key question is
that the desired E-field is expressed in the target site?
thus not only whether the applied E-field directly excites
For TMS, a high-voltage current pulse (< 300 μs), running
local neuronal elements in the target brain region but al-
through an insulated coil held tangential to the scalp, pro-
so whether it generates the specific neuronal effects re-
duces a magnetic field that painlessly penetrates the skull
quired for the chosen experimental approach (Figure 2).
and in turn produces an electric current in the underlying
Although there is some principal understanding of the
brain tissue. Importantly, the magnetic field is not atten-
neuronal effects of TMS and TCS, combining noninvasive
uated by the intermediate bone, but the induced E-field in
electrophysiological and neuroimaging techniques with
the brain simply decays exponentially with distance from
NIBS can help to verify for a specific experiment that
the TMS coil (Thielscher & Kammer, 2004). This high-
the desired neuronal effects were successfully induced.
lights the role of the scalp–cortex distance, which is
known to vary across both brain regions and individuals
and which can partially be accounted for by adjusting
Inducing Immediate (Online) Effects in Local
stimulation intensity to the actual scalp–cortex distance
Neuronal Activity
(Stokes et al., 2005). In addition, the local E-field distri-
bution depends on the anatomical distribution of brain Membrane polarization is presumably the main mecha-
tissues with different conductivities (gray matter, white nism of action for both TMS and TDC, although addi-
matter, corticospinal fluid) and the individual gyrification tional mechanisms have been discussed (Peterchev
of the underlying cortex (Opitz, Windhoff, Heidemann, et al., 2012). For TMS, the E-field dynamics are sufficiently
Turner, & Thielscher, 2011; Thielscher, Opitz, & Windhoff, fast and strong to depolarize the neuronal membrane to
2011). The E-field induced by TCS, in contrast, has to pass suprathreshold levels, presumably at the level of axons or
through the bone, which is a major barrier of low conduc- axon terminals (Aberra, Wang, Grill, & Peterchev, 2020),
tivity, causing large portions of the stimulation current to a degree that APs emerge and spread along the mem-
to be shunted via the scalp and to enter the skull via brane. These APs then transsynaptically affect connected
openings, such as eyeballs, ear canals, or small foramen neurons, causing excitatory and inhibitory postsynaptic
for the cranial nerves. Within the brain, the E-field distri- potentials, via glutamate and GABA-A/GABA-B receptors,
bution depends again on the distribution of brain tissues, depending on the initially depolarized neuron type.
but unlike for TMS, the TCS-related E-field extends across Spatial and temporal integration of postsynaptic poten-
a much larger brain volume (depending on the specific tials then causes excitatory and inhibitory postsynaptic
electrode type and montage; Opitz, Paulus, Will, Antunes, neurons to fire. Although the intracortical circuitry re-
& Thielscher, 2015; Datta et al., 2009). Importantly, E-field sponding to TMS has been studied in great detail using
simulations based on anatomically precise individual paired-pulse protocols and pharmacological interven-
head models revealed that the location of the maximum tions in the primary motor cortex (Di Lazzaro &
E-field varies across brain regions and individuals and Ziemann, 2013; Di Lazzaro, Ziemann, & Lemon, 2008)
is not simply located directly underlying the TMS coil and some insights have been generated by work in
(Weise, Numssen, Thielscher, Hartwigsen, & Knösche, 2020; rodents or nonhuman primates (Romero, Davare,
Gomez-Tames, Hamasaka, Laakso, Hirata, & Ugawa, 2018) Armendariz, & Janssen, 2019; Li et al., 2017; Mueller
or TCS electrode (Opitz et al., 2015; Saturnino, Antunes, & et al., 2014), the specific circuit motifs activated in most
Thielscher, 2015). To establish that an effective stimula- human cortical regions can only be speculated about. In
tion intensity is achieved at the target coordinate, indi- any case, TMS evokes highly synchronized neuronal re-
vidualized E-field modeling is advisable for both TCS sponses of entire intracortical circuits, not only because
(Alekseichuk, Falchier, et al., 2019; Kasten, Duecker, the E-field initially depolarizes a large number of different
Maack, Meiser, & Herrmann, 2019) and TMS (Weise et al., neurons but also because the activation spreads among
2020; Bungert, Antunes, Espenhahn, & Thielscher, 2017). them. Consequently, there will be both excitation and
Yet, although spatial E-field parameters are reliably simulated, inhibition within a neuronal population or brain region
its absolute intensity ( V/m) at the target coordinate unfor- after TMS, and the net effect on its excitability or informa-
tunately is more uncertain (Saturnino, Thielscher, Madsen, tion processing capabilities is complex (as will be dis-
Knösche, & Weise, 2019). cussed for Arrow 4).

Bergmann and Hartwigsen 7


For TCS, the E-field is much weaker and assumed to (online) in a specific brain region to demonstrate its
merely shift the membrane potential slightly toward de- causal relevance for a cognitive function (Figure 2B).
polarization or hyperpolarization, changing neuronal ex- Unfortunately, we lack a precise understanding of most
citability on a subthreshold scale (Liu et al., 2018; Stagg & neuronal activity patterns implementing a specific com-
Nitsche, 2011), either constantly (TDCS) or rhythmically putation and thus do not know exactly what to interfere
(TACS). The E-fields induced by standard stimulation in- with. Accordingly, TMS for interference often uses either
tensities (1–2 mA) in humans are much lower than those high-intensity single-pulse TMS when aiming for a good
in mice or monkeys (Alekseichuk, Mantell, Shirinpour, & temporal resolution and thus short period of interference
Opitz, 2019), and the effectiveness of TCS in humans is (Amassian et al., 1998) or short TMS burst at a high fre-
thus highly debated (Filmer, Mattingley, & Dux, 2020; quency (mostly 10–20 Hz) covering several hundred mil-
Liu et al., 2018; Vöröslakos et al., 2018), although liseconds to ensure sufficiently long disruption of
TCS-induced E-fields as small as 0.2–1 V/m have already neuronal processing (Capotosto, Babiloni, Romani, &
proven effective in causing tiny shifts in spontaneous Corbetta, 2012; Taylor, Nobre, & Rushworth, 2007).
neuronal firing rates (Krause, Vieira, Csorba, Pilly, & Although the TMS-evoked neuronal activity in local
Pack, 2019; Liu et al., 2018; Reato, Rahman, Bikson, & circuits is complex, it can safely be assumed that TMS
Parra, 2010). Although small, the TCS-induced E-field is (i) excites random neural elements (those optimally
broad, and the effects may accumulate across large neu- located relative to the E-field), including those not acti-
ron populations. Again, entire circuits will be stimulated vated by the task; (ii) results in subsequent suppression
both directly by the E-field and indirectly via synaptic con- of neuronal activity, also in neurons activated by the task,
nections, but the response will be less synchronized and for ∼50–150 msec after initial excitation (Li et al., 2017;
more strongly dependent on ongoing brain activity com- Moliadze, Zhao, Eysel, & Funke, 2003), potentially
pared to TMS. Importantly, the impact of the E-field on because of GABA-B-receptor-mediated inhibition, paral-
different neuronal structures at the cellular level depends leling the motor cortical phenomena of the cortical silent
on both their shape and their orientation with the brain period (Stetkarova & Kofler, 2013; Chen, Lozano, &
and is thus highly complex, with the net effect on a given Ashby, 1999) and long-interval intracortical inhibition
neuron depending on the integration of diverse depolar- (McDonnell, Orekhov, & Ziemann, 2006; Valls-Solé,
ization and hyperpolarization of its parts (Rahman et al., Pascual-Leone, Wassermann, & Hallett, 1992); and (iii)
2013). Even more so, the net effect on an entire brain region causes a highly synchronized neuronal activity in the tar-
arises from the integration of the individual neurons’ excit- get region based on the time-locked excitation–inhibition
ability changes, highlighting the impossibility of a simple pattern artificially evoked in a comparably large neuron
relationship between TCS polarity and the resulting net population (Romero et al., 2019). We will discuss possi-
excitability change of the target brain region. ble implications of these neuronal effects for the neuro-
nal computations mediating cognition for Arrow 4.
Although stimulation intensity and frequency are likely
Quantifying Excitability and Connectivity key parameters for determining a successful interference
protocol, there has been no systematic comparison of
The quantification of motor or visual cortical excitability
stimulation intensities, frequencies, and train durations
via TMS-induced MEPs or phosphenes comes with the in-
regarding their principal suitability for interference
herent proof of suprathreshold stimulation of target neu-
protocols.
rons (Figure 2A). There are many studies that elegantly
employ MEP or phosphene measurement to demon-
strate the modulation of motor or visual cortical excitabil- Modulating (“Gating”) Neuronal Excitability
ity under various task conditions (Lepage, Saint-Amour,
Online TDCS is often supposed to modulate the excitabil-
& Théoret, 2008; Sparing et al., 2002). Using dual-coil
ity of a certain brain region during a task with the ratio-
TMS, effective connectivity with those brain regions can
nale to facilitate task-relevant neuronal processing
be assessed in a task-dependent fashion (Murakami,
(Figure 2C). Animal work has indeed shown a polarity-
Restle, & Ziemann, 2012; Davare, Lemon, & Olivier,
dependent modulation of spontaneous neuronal spiking
2008). However, these studies typically use NIBS in a cor-
(Fröhlich & McCormick, 2010; Bindman, Lippold, &
relative manner, treating target brain region excitability as
Redfearn, 1964). However, given the complexity of neu-
a DV, not an IV, and do not probe the causal impact of
ronal excitability changes, in humans, anodal and cathodal
brain activity on cognition.
TDCS can generally not be equated with excitability in-
crease and decrease outside the primary motor cortex
(M1). This issue is somewhat resolved for transcranial
Interfering with Spontaneous or Task-related
random noise stimulation, composed of various (particu-
Neuronal Activity
larly high, > 140 Hz) TACS frequencies (Terney, Chaieb,
Many classic cognitive neuroscience TMS studies aim to Moliadze, Antal, & Paulus, 2008). Even for a fixed polarity,
interfere with neuronal activity during task processing no simple dose–response curve can be observed for TCS

8 Journal of Cognitive Neuroscience Volume X, Number Y


(Esmaeilpour et al., 2018), and these nonlinear effects entrainment per se (Vossen, Gross, & Thut, 2015; Zaehle,
complicate the determination of appropriate stimulation Rach, & Herrmann, 2010).
dosages. Irrespective of these challenges, a noteworthy
approach is to induce excitability changes during a learn-
Inducing Aftereffects (Offline) in Local Neuronal
ing task to gate learning-induced synaptic plasticity, which
Excitability Based on Synaptic Plasticity
can result in long-lasting effects, not resulting from
stimulation- but learning-related plasticity (O’Shea et al., Repetitive TMS (rTMS) or prolonged TDCS can produce
2017; Snowball et al., 2013; Vollmann et al., 2013). This transient changes in neuronal excitability, mediated by
approach effectively increases the low anatomical preci- synaptic plasticity and outlasting the stimulation protocol
sion of the TCS by the task-related activation of highly itself by minutes to hours (Figure 2D and E). Several NIBS
specific circuits. protocols have been developed for M1, producing bidirec-
tional changes in corticospinal excitability as indexed by
MEP amplitude, primarily depending on the frequency
or pattern of rTMS or the polarity of TDCS (Ziemann
Entraining Neuronal Activity (and Oscillations)
et al., 2008). In principle, classic high-frequency (∼5 Hz)
Rhythmic TMS or TACS at a certain frequency is used to versus low-frequency (∼1 Hz) rTMS results in long-term
entrain neuronal activity with the aim to synchronize and potentiation (LTP)-like and long-term depression (LTD)-
enhance endogenous brain oscillations (Figure 2C) and test like facilitation and inhibition of corticospinal excitability,
their causal role for cognition (Vosskuhl et al., 2018; Antal respectively (Fitzgerald, Fountain, & Daskalakis, 2006),
& Herrmann, 2016; Herrmann, Rach, Neuling, & Strüber, whereas for theta burst stimulation, the specific timing
2013; Thut, Schyns, & Gross, 2011; Thut, Veniero, et al., of TMS trains and pauses determines the direction of ef-
2011). However, the underlying neurophysiological fects (Huang, Edwards, Rounis, Bhatia, & Rothwell,
assumptions are often not made explicit. Neuronal oscil- 2005). Likewise, TDCS in classic M1-contralateral forehead
lations in EEG/magnetoencephalography (MEG) reflect montage produces lasting increases and decreases in cor-
the summed potentials/fields from large synchronized ticospinal excitability, respectively, depending on whether
neuron populations with parallelly oriented dendritic anode or cathode overlays M1 (Nitsche & Paulus, 2000).
trees (Cohen, 2017). Their amplitude increases when Stimulation intensity and duration are crucial determi-
the postsynaptic activity of more neurons becomes syn- nants for offline effects with both TMS (Ziemann et al.,
chronized, and the entrainment of a neuronal oscillation 2008) and TCS (Nitsche et al., 2008). However, Bonaiuto
by rhythmic NIBS typically refers to the synchronization and Bestmann (2015) emphasized that the “sliding-scale
of spontaneously but yet independently oscillating rationale” (assuming the magnitude of cortical excitability
neurons (Thut, Schyns, et al., 2011). However, it is also increases to scale with stimulation intensity) is incorrect,
possible that random neuronal activity is entrained in- as nonlinearity has been clearly demonstrated even for
stead of an already ongoing endogenous oscillation M1 (Batsikadze, Moliadze, Paulus, Kuo, & Nitsche, 2013;
(Herring, Esterer, Marshall, Jensen, & Bergmann, 2019). Moliadze, Atalay, Antal, & Paulus, 2012). For a detailed dis-
Entrainment may also work differently for TMS and TCS. cussion of the neurophysiological mechanisms mediating
Whereas rhythmic (suprathreshold) TMS may directly NIBS-induced LTP/LTD-like plasticity, the reader is re-
evoke waves of synchronized excitation and inhibition, ferred to previous reviews (Hoogendam, Ramakers, & Di
potentially phase-resetting existing oscillatory activity Lazzaro, 2010; Ziemann et al., 2008). Importantly, these af-
(Herring, Thut, Jensen, & Bergmann, 2015), TACS merely tereffects show large intraindividual and interindividual
shifts the membrane potential forth and back, biasing variability (as discussed below), often emerge with a cer-
spontaneous neuronal firing. The weaker impact of tain delay (Huang et al., 2005), and wash out after an un-
TACS thus likely requires it to be more well targeted, known duration, typically 30–60 min (Ziemann et al.,
for example, adjusted to the individual frequency of 2008). Although the effectiveness of NIBS can immediately
the target oscillation ( Vosskuhl et al., 2018; Thut et al., be assessed via MEP amplitudes for M1, such a manipula-
2017). Unfortunately, direct proof of neuronal entrain- tion check for other brain regions requires neuroimaging
ment during rhythmic NIBS is currently difficult to techniques (see below). Although common practice, we
impossible because of the strong NIBS-related artifacts cannot assume every NIBS protocol to easily translate from
(Gebodh et al., 2019; Rogasch et al., 2017; Noury, motor to nonmotor regions, and without a manipulation
Hipp, & Siegel, 2016; Herring et al., 2015; Ilmoniemi & check, we can only hope for the desired excitability effects
Kičić, 2010). However, there have been a few successful to occur in the target region.
attempts using more indirect measures of neuronal
entrainment (Herring et al., 2019; Hanslmayr, Matuschek,
Mapping NIBS-related Neuronal Effects
& Fellner, 2014; Helfrich, Schneider, et al., 2014; Thut,
with Neuroimaging
Veniero, et al., 2011). Note that any lasting increase in oscil-
latory power after TACS reflects synaptic aftereffects (off- Both online and offline NIBS effects can be assessed in
line) in the oscillation-generating circuits, not ongoing humans with noninvasive neuroimaging techniques

Bergmann and Hartwigsen 9


(e.g., fMRI, EEG, or MEG). For a detailed discussion of the Paulus, & Nitsche, 2012) TCS–fMRI studies reported
challenges associated with the combination of NIBS and neu- widespread BOLD effects (Turi, Paulus, & Antal, 2012),
roimaging, see previous review articles (Bergmann et al., although for TCS, direct unfocal stimulation effects (i.e.,
2016; Bestmann & Feredoes, 2013; Siebner, Bergmann, direct effects of the widespread E-field on neuronal activ-
et al., 2009). Neuroimaging is crucial to provide proof of ity outside the target region) are difficult to disentangle
target engagement, that is, to verify the assumption that from actual network effects (i.e., spread of local E-field-
the applied NIBS protocol has effectively induced the in- induced changes in neuronal activity to remote regions
tended neuronal activity in the target region. The high via long-range axonal projections and synaptic connec-
spatial resolution of BOLD fMRI helps to detect net tions). In addition, concurrent TMS–EEG studies report
changes in spontaneous or task-related neuronal activity TMS-evoked potentials spreading within the targeted net-
via resting-state or task fMRI, but for demonstrating work (Harquel et al., 2016; Massimini et al., 2005), and
entrainment or interference effects, the superior tempo- dual-coil TMS studies typically build on this feature when
ral resolution of EEG or MEG is usually required. testing effective connectivity between two brain regions
Neuroimaging also allows to screen for unintended coac- (Silvanto, Lavie, & Walsh, 2005; Ferbert et al., 1992).
tivation of nontarget brain regions, which may otherwise Neuroimaging can be used to read out both immediate
cause confounding and prevent the unambiguous identi- (online) effects as well as subsequent (offline) effects me-
fication of structure–function relationships (cf. Arrows 3 diated via synaptic plasticity (Bergmann et al., 2016). For
and 4). In the absence of behavioral effects, neuronal aftereffects on remote neuronal activity, the question re-
activity may be the only readout available to investigate mains, however, whether they are caused by local synap-
network effects such as compensation (cf. Arrow 3), tic plasticity in the target site, subsequently affecting
whereas in the presence of behavioral effects, brain– remote activity via changes in functional connectivity or
behavior correlations may further corroborate the causal via synaptic plasticity in the remote site itself induced by
link between NIBS-induced neuronal and behavioral spread of activity during the stimulation.
effects in terms of dose–response relationships.
Consequences for Network Activity
FROM LOCAL TO NETWORK EFFECTS
In any case, both online and offline effects on remote
(ARROW 3)
nodes can be functionally relevant. For instance, in a
The third cause–effect pair (Arrow 3) refers to the impact of consecutive TMS–fMRI study, the rTMS-induced increase
NIBS-induced local neuronal activity on other connected in the inhibitory influence of the stimulated area on a
nodes of the target network. This spread of activation remote node predicted the individual TMS-induced
may be desired or considered a potential confound, but response delay in a language task (Hartwigsen et al.,
it is in any case an inherent feature of the brain, not a 2017). Inhibitory stimulation effects are thus not restricted
shortcoming of the method. Network effects always need to the stimulated area but can affect large parts of the
to be considered when attributing changes in cognitive network, also modulating the functional interaction of
function to NIBS-induced changes in the targeted brain its elements. Such remote network effects remain hidden
region. in purely behavioral studies if single-site TMS or TCS is used
and are usually ignored when drawing conclusions about
the causal relevance of the stimulated area for a given
Remote Effects of NIBS
task. Yet, network effects are potential confounders, es-
The most direct evidence of transsynaptic spread is the pecially when relying on plasticity-inducing offline proto-
MEP after TMS of M1, which relies on several synaptic cols that leave the brain time for adaptive plasticity in
connections from the initially excited neural elements response to the intervention and rapid short-term reorga-
in M1 via corticospinal output neurons and spinal moto- nization of the network. Note that remote effects can be
neurons to the muscle. Yet, cortico-cortical spread has inhibitory, facilitatory, or both in different parts of the
been demonstrated by combined NIBS–fMRI studies, network, and the direction of network effects is difficult
which revealed strong remote effects of TMS in networks to predict a priori. NIBS-induced inhibition of a key target
for various motor and cognitive functions (Bergmann node sometimes decreases task-related activity in larger
et al., 2016; Bestmann & Feredoes, 2013). Concurrent parts of the network (Hartwigsen et al., 2017; O’Shea,
TMS–fMRI studies, applying TMS to M1, FEFs, or intrapar- Johansen-Berg, Trief, Göbel, & Rushworth, 2007), which
ietal sulcus, found strong (dose- and state-dependent) in turn disinhibits and increases activity in other network
effects in remote but anatomically connected cortical nodes compensating for the disruption and prevention of
and subcortical areas (Ruff et al., 2008; Bestmann, behavioral effects. Such compensatory upregulation can
Baudewig, Siebner, Rothwell, & Frahm, 2003, 2005), even occur in contralateral homologous regions (O’Shea
for subthreshold intensities (Bestmann et al., 2003). et al., 2007), ipsilateral network nodes (Hallam,
Likewise, concurrent (Antal, Polania, Schmidt-Samoa, Whitney, Hymers, Gouws, & Jefferies, 2016), and neigh-
Dechent, & Paulus, 2011) and consecutive (Polania, boring regions relevant for other cognitive functions,

10 Journal of Cognitive Neuroscience Volume X, Number Y


including domain-general areas (Hartwigsen et al., 2017). in task performance are expected. Yet, the term “virtual
This short-term reorganization in response to focal dis- lesion” is misleading, because TMS does not simply
ruption stresses the strong potential for flexible redistri- switch off a brain region, and offline and online ap-
bution of resources and the high degree of degeneracy proaches rely on different neuronal mechanisms (cf.
in the brain (Price & Friston, 2002; Edelman & Gally, Arrow 2). Moreover, “virtual lesions” can only explain per-
2001). Combining NIBS with neuroimaging provides a formance impairments, and improvements in response
means of mapping both local and remote network effects to “inhibitory” protocols are often referred to as “para-
at the systems level and relating these effects to changes doxical facilitation” ( Walsh & Cowey, 2000; Kapur,
in behavior. 1996). The alternative rationale for demonstrating causal
relevance of a brain region or neuronal activity pattern
for a given cognitive function is to facilitate it during task
FROM NEURONAL (NETWORK) ACTIVITY TO
execution and show positive consequences for perfor-
COGNITIVE EFFECTS (ARROW 4)
mance. Again, this is typically tried either with online
The fourth cause–effect pair (Arrow 4) refers to the tran- modulation (Figure 2C), entraining task-relevant oscilla-
sition from neuronal network activity to a cognitive ef- tions via rhythmic TMS or TACS or increasing immediate
fect. The latter is not directly observable but must be cortical excitability via TDCS during a task, or by offline
operationalized as a specific task to be assessed via be- facilitation (Figure 2D). The latter is supposed to induce
havioral performance. Importantly, there is no one-to- a lasting increase of spontaneous neuronal activity during
one mapping of brain activity to cognitive functions, as a subsequent task via the strengthening of synapses via
the same region is likely involved in multiple functions facilitatory offline NIBS. Because it is generally easier to
and the same cognitive function relies on the interaction disturb than to improve an insufficiently understood pro-
of multiple regions. Moreover, NIBS protocols can influ- cess, the “facilitatory” approach is used less frequently.
ence the interaction between network activity and cogni- NIBS-induced facilitation of behavior is nonetheless
tive function but rarely produce a direct behavioral tempting, as it opens interesting avenues for therapeutic
output. The main question of this section is thus whether applications or neuroenhancement. Importantly, the
the desired modulation of local and network activity af- choice of NIBS protocols for a specific study is often built
fects the target cognitive function of interest it is assumed on oversimplified assumptions, which partially explains
to mediate. Both online and offline NIBS approaches can the many null findings and controversial results. Below,
either facilitate or inhibit a cognitive function (Figure 1B), we will discuss the possible mechanisms of action trans-
and although a causal discovery per se (e.g., Cortical lating neuronal into cognitive effects and highlight some
Area X but not Y is causally relevant for a Cognitive of the modulating factors.
Process A but not B) can be made independently from
the estimation of direction, size, or specific function of
Impairing a Cognitive Function by Online TMS
the cause–effect relationship, the latter is crucial for
understanding the neuronal mechanism underlying a The most effective approach for this aim may be online
cognitive function and for developing theory-based applica- interference via TMS (Figure 2B), which presumably
tions. It can thus be considered a key challenge is to pre- builds on three neuronal effects (cf. Arrow 2). First, the
dict a priori the direction and size of the induced effects initial excitation of random neural elements causes neu-
for a given NIBS protocol, cognitive function, and exper- ronal noise in the stimulated circuits (Ruzzoli, Marzi, &
imental setting. Miniussi, 2010; Siebner, Hartwigsen, Kassuba, &
Rothwell, 2009). Noise pervades all levels of information
processing in the nervous system, from receptor signal
Impairing vs. Improving Cognitive Functions
transduction to behavioral responses (Faisal, Selen, &
with NIBS
Wolpert, 2008). The artificial induction of noise may im-
Many NIBS studies in cognitive neuroscience rely on the pair or delay task-relevant neuronal computations be-
“virtual lesion” approach to map causal relationships be- cause neural activity needs to be sampled longer to
tween neuronal activity in a given brain region and a cog- discriminate signal and noise. Second, the initial excita-
nitive function of interest, assuming that disturbing or tion is inevitably followed by GABA-B-ergic feedback inhi-
inhibiting task-related neuronal activity by an online or bition, suppressing neuronal activity for ∼50–150 msec
offline NIBS protocol will result in impairment of the in- after TMS (Inghilleri, Berardelli, Cruccu, & Manfredi,
vestigated cognitive function. The online interference ap- 1993; Haug, Schönle, Knobloch, & Köhne, 1992), inter-
proach (Figure 2B) aims at transiently disturbing a rupting and delaying neuronal processing or even caus-
cognitive function with TMS during task execution, ing signal loss during crucial processing steps. This
whereas the offline inhibition approach (Figure 2E) relies effect may come closest to the “virtual lesion” idea of
on a decrease in cortical excitability during the task, me- silencing neuronal activity. Third, the evoked excitation–
diated by the preceding weakening of synapses after an inhibition sequence artificially synchronizes larger neu-
inhibitory offline NIBS protocol. In both cases, decreases ron populations, thereby lowering the number of

Bergmann and Hartwigsen 11


possible neuronal activity patterns in the network. This activity (Romei et al., 2016; Thut, Veniero, et al., 2011),
loss of entropy (Shannon & Weaver, 1949) in local neu- and whether a TMS burst impairs or improves a cognitive
ronal activity reduces the information representation ca- function may thus depend on whether or not rhythmi-
pacity of the synchronized network (Hanslmayr, Staudigl, cally synchronized brain activity in the target network is
& Fellner, 2012; Schneidman et al., 2011; Tononi, 2008), beneficial for the task.
leading to a degradation of task-relevant information and
a disruption of neuronal computations. This may result in
Bidirectionally Modulating (Gating) a Cognitive
prolonged processing time (because of the need for
Function with Online or Offline NIBS
compensatory iterations or the recruitment of additional
processing resources) or even an incorrect outcome of In contrast to disrupting or actively driving the neuronal
the computation. Importantly, an online disruption does processes mediating a cognitive function, tonic online
not leave the targeted network time for functional reor- modulation via TDCS (Figure 2C) is assumedly able to bi-
ganization (see below), leading to stronger stimulation directionally modulate (decrease or increase) task-related
effects and simplifying the conclusions in comparison neuronal activity, depending on stimulation polarity
to offline approaches. (among other factors). As discussed for Arrow 2, there
is no straightforward relationship between TDCS polarity,
intensity, and net excitability changes. There is also no
Improving a Cognitive Function via Entrainment by
simple mapping from cortical excitability to cognitive
Online TACS or TMS
performance for any given NIBS protocol. It is well con-
To actively improve task-related function, specific as- ceivable that a net excitability increase in the circuits me-
sumptions are needed regarding the neuronal mecha- diating task-relevant computations augments task-related
nisms of action. Comparably simple targets are signal and thus boosts signal-to-noise ratio (SNR).
neuronal oscillations, supposedly underlying a variety of Alternatively, increased excitability may also augment
cognitive functions (Buzsáki & Draguhn, 2004). To en- spontaneous but task-irrelevant activity, thereby increas-
train a neuronal oscillation by means of phasic online ing noise and lowering the SNR. The same ambiguity ap-
modulation via rhythmic TMS or TACS (Figure 2C), not plies also for excitability decreases. Importantly, although
only the brain region or network but also the oscillatory additional noise may degrade task-relevant SNR for a
frequency needs to be made explicit. When able to tran- well-tuned neuronal representation (e.g., when distin-
scranially entrain (and augment) a neuronal oscillation guishing similar stimuli in a discrimination task) and thus
and thereby improve the cognitive function, this provides impair task performance, noise may be beneficial for
proof of its causal relevance ( Vosskuhl et al., 2018). other tasks (e.g., by lifting a weak perceptual stimulus
However, whether an increase in synchronization is ben- above threshold in a detection paradigm). The actual
eficial depends on the very mechanism by which the os- effects of online NIBS depend on the complex interaction
cillation mediates the relevant neuronal computations. of spontaneous and task-induced brain states, specific
Beyond mere excitability fluctuations (Bergmann, Lieb, task demands, participant-specific characteristics, and
Zrenner, & Ziemann, 2019; Bergmann et al., 2012; stimulation parameters (Fertonani & Miniussi, 2017).
Schroeder & Lakatos, 2009), oscillations supposedly en- Similar considerations also apply to the rationale of in-
able more complex processes, such as interarea (Fries, hibition or facilitation via offline TMS or TCS protocols
2015) and cross-frequency communication via phase– (Figure 2D and E), although the bidirectional modulation
phase or phase–amplitude coupling ( Jensen & Colgin, of cortical excitability after offline NIBS is based on differ-
2007), phase coding ( Jensen, Gips, Bergmann, & ent neuronal mechanisms (cf. Arrow 2). The LTP-like
Bonnefond, 2014; Lisman & Jensen, 2013), and poten- strengthening or LTD-like weakening of synapses in the
tially phase-dependent plasticity (Bergmann & Born, 2018). target network results only indirectly in subsequent in-
These complex processes are more difficult to optimize, creases or decreases in neuronal excitability and respec-
although (multifocal) TACS has successfully been used to tive changes in spontaneous and task-related neuronal
produce behaviorally relevant interarea synchronization activity. It should also be noted that evidence for a bidir-
(Reinhart & Nguyen, 2019) and working memory en- ectionality of offline effects is based almost entirely on
hancement (Alekseichuk, Turi, Amador de Lara, Antal, & the respective modulation of MEP amplitudes after M1
Paulus, 2016). However, an increase in synchronization, stimulation and may not easily generalize to other mon-
which may unintentionally also recruit task-irrelevant tages and cortical areas (Parkin, Bhandari, Glen, & Walsh,
neurons, or an entrainment-induced phase shift of the 2019). Importantly, as for the effects of online modula-
endogenous oscillation may also be detrimental to task- tion, excitability changes are not homogenously distrib-
relevant neuronal computations. Interestingly, the same uted within the targeted brain circuits, because only a
TMS protocols (e.g., four to five pulse bursts of 10- or random selection of functionally heterogeneous synapses
20-Hz TMS) are often used for both behavioral inter- is affected. Despite a possible net facilitation or inhibition
ference (Hartwigsen, Price, et al., 2010) and neuronal of excitability, random changes in either direction are
entrainment, for example, of alpha or beta oscillatory most likely to produce noise in the neuronal activity

12 Journal of Cognitive Neuroscience Volume X, Number Y


patterns generated by these circuits, with the above- remain unknown, and the above explanations are mainly
discussed positive or negative consequences for task per- used in a post hoc fashion. It is even conceivable that, in
formance. These considerations are highly relevant for some cases, for example, as a consequence of increased
training studies and therapeutic applications of NIBS, neuronal noise, the true direction of the NIBS effect on
which often assume an offline NIBS-induced facilitation cognitive performance varies across participants or even
to improve behavior in a subsequently performed task within participants across trials. In such a case, the mere
(Luber & Lisanby, 2014). increase in variance (beyond measurement noise) after
NIBS may be considered evidence for a cause–effect re-
lationship, even when lacking a clear direction. However,
The Paradox of Paradoxical Facilitation
such a relationship would be less informative regarding
“Paradoxical facilitation” usually refers to an unexpected the neuronal mechanisms underlying the cognitive
positive effect of an “interfering” or “inhibitory” NIBS function and more difficult to exploit for therapeutic
protocol on a cognitive function. As emphasized above, applications.
noise is a central concept to explain the cognitive effects Various NIBS-induced neuronal effects can affect task-
of NIBS protocols (Ruzzoli et al., 2010; Siebner, relevant neuronal computations, although little is known
Hartwigsen, et al., 2009). Taking stochastic resonance about the factors making a specific neuronal process sus-
into account, adding noise to a nonlinear system like ceptible to or robust against this influence. NIBS effects
the human brain may produce opposite effects. Whereas are often small, and the brain is capable to compensate
an appropriate amount of noise can add to the weak for weak disturbances, likely contributing to the null ef-
neuronal signal of a subthreshold stimulus, elevate it fects observed in many NIBS studies.
above threshold, and result in behavioral facilitation
(Schwarzkopf, Silvanto, & Rees, 2011; Miniussi, Ruzzoli,
Multi-site Approaches to Study
& Walsh, 2010), exceeding noise levels may rather mask
Network Interactions
the task-relevant neuronal signal. Importantly, the NIBS-
induced activity or neural noise is not totally random Because all cognitive functions rely on distributed pro-
(Ruzzoli et al., 2010) and also not independent of the cesses organized in large-scale neural networks, there is
task-induced neural activity or brain state. Thus, depend- increasing interest in disrupting several network nodes
ing on the activated neuron population, the induced ac- for a given function in a simultaneous or subsequent
tivity can even be considered both as noise and as part fashion to study stimulation-induced network effects on
of the signal (Miniussi et al., 2010). If the induced neuronal cognitive functions. Multifocal TMS can provide insights
noise is synchronized with the ongoing relevant activity into functional network interactions and elucidate their
(Ermentrout, Galán, & Urban, 2008), it may augment compensatory potential. Functional interactions can
the signal (Miniussi, Harris, & Ruzzoli, 2013). In other be studied either online by simultaneously targeting
words, behavioral facilitation may result from an opti- more than one area (“multi-site” approach) or by com-
mum level of noise in the system. Although originally de- bining offline and online TMS over different regions
scribed for online TMS studies (Abrahamyan, Clifford, (“condition-and-perturb” approach). Multi-site TMS ap-
Arabzadeh, & Harris, 2011), these principles also hold proaches are particularly suited to map the immediate
for offline TMS and TCS protocols (cf. Fertonani & consequences of the disruption of several brain regions
Miniussi, 2017). Indeed, offline NIBS applied before task because the acute TMS-induced interference during
processing may transiently prime activity in the stimulated task performance leaves the system no time to develop
area to a level that facilitates subsequent task perfor- adaptive plasticity. This allows to test whether the inter-
mance, although homeostatic metaplasticity may lead to ference effect over one area may be increased by the si-
opposite effects (see below). Besides positive conse- multaneous disruption of other key regions, ipsilateral
quences of noise, paradoxical facilitation may also arise (Ellison & Cowey, 2009) or contralateral (Hartwigsen,
from NIBS-induced inhibition of task-irrelevant areas that Baumgaertner, et al., 2010) to the stimulation site. In a
compete for resources (“addition-by-subtraction” [Luber complementary fashion, the condition-and-perturb ap-
& Lisanby, 2014]) or the disruption of distracting stimulus proach can be used to study rapid network redistribution
elements, facilitating task-relevant processing ( Walsh, and compensation (Hartwigsen, 2018). It combines the
Ellison, Battelli, & Cowey, 1998). Finally, stimulation- plastic aftereffects of offline modulation with the imme-
induced disinhibition of distant connected areas may fa- diate perturbation effects of online interference, follow-
cilitate cognitive processing (Sandrini, Umiltà, & Rusconi, ing the rationale that offline conditioning of one area
2011). may sensitize another network node to the disruptive ef-
Unfortunately, given the complex interaction between fect of online interference (e.g., Hartwigsen et al., 2012).
task, brain state, stimulation protocol, and intensity, in In some cases, offline conditioning of a single target area
most cognitive neuroscience experiments, the exact cir- does not affect task-related behavior, whereas additional
cumstances under which a given NIBS protocol results, online disruption of a second area effectively impairs task
on average, in behavioral impairment or facilitation performance, unmasking the disruptive effect of the

Bergmann and Hartwigsen 13


offline protocol (Hartwigsen et al., 2015; Sack et al., effects of NIBS-induced modulations of network dynamics
2009). A likely explanation is that both areas contribute on behavioral measures. They rely on the assumption
to the function of interest and that offline conditioning that perturbation of the dynamics of a biophysical net-
of one area can be compensated by a stronger contribu- work model via membrane depolarization affects cogni-
tion of the other node, changing their functional weights tive function (e.g., value-based decision-making) in a
within the network. The additional online perturbation predictable way, by modulating the susceptibility of net-
increases the overall “lesion load” and thus results in cog- work dynamics to background noise. Thereby, they allow
nitive disruption. When combined with neuroimaging, for predicting large-scale network effects of neurostimula-
condition-and-perturb approaches can also be used to tion that can be experimentally validated (Bonaiuto et al.,
study rapid reorganization at the network level. O’Shea 2016; Hämmerer et al., 2016).
et al. (2007) demonstrated that offline TMS over the left
premotor cortex decreased task-related activity in the
FROM COGNITIVE FUNCTION TO
stimulated area during an action selection task and in-
BEHAVIORAL RESPONSE (ARROW 5)
duced compensatory upregulation in other areas of the
motor network, including the contralateral homologous The fifth cause–effect pair (Arrow 5) refers to the impact
region. Targeting the “reorganized” homologous premo- of a cognitive function (or its modulation) on the perfor-
tor cortex with subsequent online TMS impaired task per- mance in a specific behavioral task. The cognitive func-
formance, demonstrating the functional relevance of the tion of interest needs to be operationalized by a
observed compensatory upregulation. This shows how specific task to measure how it is affected by a given
the combination of neuroimaging and multifocal TMS NIBS protocol, and this task has to be sufficiently difficult
can provide insight into the compensatory dynamics of to be sensitive enough for the (usually very small) cogni-
task-specific neural networks. tive changes induced by NIBS. Usually, the interaction of
several cognitive functions is necessary for task comple-
tion, and a control task is needed that differs selectively
Cognitive Models and
in the target cognitive component to contrast out the in-
Computational Neurostimulation
fluence of other cognitive processes and establish task
To map stimulation-induced changes on a cognitive func- specificity.
tion of interest, a valid cognitive model is mandatory that
can be translated into a task. For instance, dual-route
The Benefit of Behavioral Modeling Approaches
models have been used to explain stimulus–response
compatibility effects in conflict tasks (Ridderinkhof, To bridge the gap between cognitive model and behav-
2002a). Such models assume parallel routes for ioral outcome measures, distributional analyses can
decision-making that can be dissociated behaviorally provide insight into different response strategies
(i.e., a direct, stimulus-based activation route and a con- (Ridderinkhof, 2002a). One advantage relative to com-
trolled, deliberate response activation route) and con- posite measures like mean response speed or task accu-
verge at the level of response activation processes. The racy is that they take the whole response distribution
dynamics of these processes can be captured with distri- into account (i.e., both correct and incorrect responses)
butional analyses that map interference effects during and are more sensitive to experimental dynamics and
decision-making (see next section). Thereby, cognitive individual differences in response strategies (van den
models help to decompose abstract and complex con- Wildenberg et al., 2010). They help to overcome the poor
structs into several subcomponents that can be opera- statistical sensitivity of composite measures ( Voss,
tionalized by specific tasks. NIBS can probe the Nagler, & Lerche, 2013). Distributional analyses map re-
dynamics between these subcomponents and the func- sponse strategies, such as the speed–accuracy trade-off,
tional relevance of different brain regions for these by disentangling whether NIBS-induced interference in-
processes. More recently, cognitive models have been creases errors for fast responses, indicating a potential
complemented by computational neurostimulation ap- emphasis on speed, or rather decreases the overall un-
proaches, which simulate emergent network dynamics certainty in task processing, resulting in increased errors
and compare them with real data (Bonaiuto & Bestmann, for slower responses. Such analyses further help to distin-
2015). Computational models may, for example, more guish subprocesses of cognitive theories. For instance,
accurately reflect choice dynamics by modeling influ- within the framework of dual-process models, they
ences of previous trials to capture choice repetition biases have been used to dissociate the role of direct response
(Bonaiuto, de Berker, & Bestmann, 2016; Hämmerer, activation (based on a target stimulus) and selective
Bonaiuto, Klein-Flügge, Bikson, & Bestmann, 2016). As suppression of activation based on precues in conflict
humans tend to repeat recent choices in real-life situa- tasks (Ridderinkhof, 2002b). These processing dynamics
tions, modeling the choice history provides valid estima- are usually lost when relying on composite scores.
tions of decision-making outside the laboratory (Bonaiuto Distributional analyses have demonstrated TDCS-
et al., 2016). Such models generate predictions about the induced modulations of different error types underlying

14 Journal of Cognitive Neuroscience Volume X, Number Y


impulsive responses (Spieser, van den Wildenberg, instance, strong interference may suppress a visual stim-
Hasbroucq, Ridderinkhof, & Burle, 2015), as well as ulus below the perception threshold (Amassian et al.,
TMS-induced changes in response strategies during ac- 1998), resulting in decreased task accuracy that cannot
tion reprogramming (Hartwigsen & Siebner, 2015; be compensated by increased response speed, and weak
Hartwigsen et al., 2012) and conflict paradigms (van visual stimuli closer to the perception threshold might be
Campen, Kunert, van den Wildenberg, & Ridderinkhof, affected first. In contrast, slight modulations of task-
2018). Other behavioral modeling approaches rely on related activity may selectively delay response speed,
sequential sampling models such as the drift diffusion with increases in latencies preventing effects on task ac-
model (Ratcliff, Smith, Brown, & McKoon, 2016; Ratcliff, curacy. The timing of the pulses and the stimulation fre-
1978), which assumedly reflect the underlying processes quency likely play a crucial role for the outcome of a
contributing to a particular response distribution and also stimulation protocol. For instance, the absence of any im-
capture decision biases. Such models are particularly sen- pairments in task accuracy in the presence of strong de-
sitive toward slight adaptations of response strategies that lays in response speed during a visual discrimination task
may be overlooked when relying on composite mea- was explained with the employed 10-Hz rTMS protocol,
sures (Ratcliff & McKoon, 2008; Voss & Voss, 2007), espe- arguing that TMS might disrupt processing for a brief pe-
cially in NIBS studies suffering from relatively small effect riod within each 100-msec interpulse interval but never
sizes (see also Hartwigsen et al., 2015). Aside from binary completely interferes with the relevant information for
choice tasks, sequential sampling models have been discrimination (Ellison & Cowey, 2009). Rather, 10-Hz
adapted for more complex multichoice decisions (Kohl, rTMS may merely delay processing by the summed pe-
Spieser, Forster, Bestmann, & Yarrow, 2019) that might riods of disruption. A similar reasoning likely explains
better match real-life decisions. why a 500-msec stimulation period of 10-Hz rTMS does
not result in a 500-msec increase in response speed
( Walsh & Cowey, 2000). Yet, to the best of our knowl-
Composite Measures to Quantify NIBS Effects
edge, no study has systematically varied the stimulation
Yet, most NIBS studies rely on composite measures de- frequency to investigate whether higher frequencies
rived from the individual mean response speed or accu- might affect both discrimination speed and accuracy.
racy, which are usually analyzed with ANOVAs or t tests. Most studies assume that TMS might affect both task
However, NIBS studies can benefit from mixed models, speed and accuracy, but it remains unclear whether the
allowing to model nonlinear individual characteristics modulation of either parameter relies on different neuro-
and providing more flexibility when handling missing nal mechanisms. Cognitive and computational models
data (Kaarre et al., 2018; Payne & Tainturier, 2018). Such may help to specify the expected effects on both out-
approaches are especially useful for longitudinal NIBS de- come parameters a priori, as the conceptualization of
signs where missing data for single time points might different subprocesses of a cognitive function helps to
otherwise lead to participant exclusion. Sometimes, mo- dissociate the expected outcomes.
tor responses can also be assessed by electrophysiologi-
cal means, for example, for MEPs recorded from orofacial
CONFOUNDING FACTORS CHALLENGING
muscles during stimulation of motor and premotor areas
THE ASSUMED CHAIN OF CAUSATION
in speech production tasks (Möttönen, van de Ven, &
Watkins, 2014; Murakami, Restle, & Ziemann, 2011). As outlined in Figure 1B, several factors may confound
For other tasks, psychometric functions (Zazio, the hypothesized chain of causation (red boxes), by af-
Bortoletto, Ruzzoli, Miniussi, & Veniero, 2019; Cattaneo fecting both sides of the investigated core cause–effect
et al., 2011) or response biases (Riddle, Hwang, Cellier, pair, that is, the structure–function (or brain–behavior)
Dhanani, & D’Esposito, 2019; Smalle, Rogers, & relationship, namely, the targeted neuronal activity (sup-
Möttönen, 2015) may be the measure of choice. No mat- posedly caused by NIBS) and the cognitive function of
ter which behavioral measure is used, NIBS protocols are interest (supposedly causing behavioral task perfor-
likely to first affect task efficiency, leading to increased mance). Some factors are associated with the application
(or decreased) response latencies, because response of NIBS, such as the unintended costimulation of nontar-
speed is a more sensitive performance metric than task get brain regions, either directly (blue boxes) or via pe-
accuracy (Bonaiuto et al., 2016). However, sometimes, ripheral sensory pathways (green boxes). Other factors
task accuracy is affected without any influences on task (yellow box) do not originate from NIBS application
efficiency (e.g., Ward et al., 2010; Amassian et al., but rather from the experimental setup (e.g., task de-
1998). As noted above, speed and accuracy reflect differ- mands), the participants’ predisposition (e.g., current
ent cognitive strategies that can be disentangled with brain state, cognitive abilities, beliefs, and expectations),
cognitive models. With respect to the potential mecha- or an interaction of both (e.g., learning effects). All these
nisms related to TMS-induced interference effects on ei- factors either directly influence brain activity and cogni-
ther of the two processes, the severity of the interference tive function or via a modulation of their response to
effect likely depends on the perceptional threshold. For NIBS. Showing considerable intraindividual and/or

Bergmann and Hartwigsen 15


interindividual variability, unsystematic variance in these Sack, 2015). Therefore, both local and remote network
factors can compromise the overall effectiveness of a effects can confound the effect of NIBS-induced target ac-
NIBS protocol to modulate the target cognitive function tivity on the target cognitive function. For example, when
(producing false negatives), whereas systematic variance targeting higher cortical association areas like the inferior
across experimental conditions can introduce systematic parietal cortex, which integrates information from several
confounding (false positives) for both within- and modalities (Seghier, 2013), its NIBS-related activation
between-participants designs. To prevent confounding, could result from transcranial or sensory stimulation.
these factors either need to be eliminated, kept constant Complex domain-specific cognitive functions usually en-
across experimental conditions, or explicitly included as gage several domain-general processes like attention,
experimental control condition (see next section). working memory, or executive functions, and sensory dis-
ruption of these processes (or other cognitive functions
of no interest) may severely affect task processing and
Costimulation of Nontarget Regions and Networks
modulate response speed or accuracy.
The application of TMS and TCS can have side effects
creating relevant confounding of Arrow 2, namely, an
Participants’ Beliefs and Expectations
effective E-field in nontarget brain regions and the
costimulation of peripheral neuronal structures. If the Participants in NIBS studies have certain expectations and
stimulation intensity is sufficiently large, the TMS- beliefs about the impact of neurostimulation and poten-
induced E-field will reach effective levels also in adjacent tial side effects. These expectations and beliefs can influ-
no-target locations (blue boxes), especially in more ence their performance in a way that may be congruent
superficially located ones. It is thus unlikely to exclu- or incongruent with the experiment’s underlying hypoth-
sively stimulate a coordinate deep in the sulcus, limiting esis. Participants may also develop expectations and actual
focal stimulation to the gyral crowns (Siebner, 2020; knowledge with regard to the temporal structure of the
Thielscher et al., 2011). For TCS, the widespread E-field experiment and the current stimulation condition (effec-
results in even more extended off-target stimulation at tive or sham), such that they may anticipate and prepare
similar or higher intensities compared to the target. for stimulation trials in online NIBS studies or behave dif-
Although the E-field is more focal for multielectrode ferently after an offline NIBS protocol.
montages with small electrodes as compared to classical
two-electrode montages with large electrodes (Datta
Task Demands, Learning Effects, and
et al., 2009), the effective E-field will not be confined
Cognitive Abilities
by the borders of the targeted brain area. In addition,
both TMS and TCS also induce effective E-fields in the Many cognitive functions vary considerably over time
cranial periphery (green boxes), with magnitudes in the (e.g., as a consequence of learning or fatigue) and across
skin being inevitably larger than those in the brain participants. Differences in baseline performance levels
(Asamoah, Khatoun, & McLaughlin, 2019), exciting because of different cognitive abilities (low vs. high per-
efferent fibers of the facial nerve innervating the facial formers) contribute to the large interindividual variability
muscles (Chen, Chauvette, Skorheim, Timofeev, & observed in NIBS studies of cognition. In visual priming
Bazhenov, 2012), or afferent fibers of the trigeminal studies, single-pulse TMS facilitated behavioral responses
nerve innervating the scalp, face, and meninges in low performers but delayed response speed in high
(Siebner, Auer, Roeck, & Conrad, 1999; Schmid, Møller, performers, and these effects interacted with task diffi-
& Schmid, 1995). In particular, the TCS-induced E-field, culty, indicating a complex interaction between stimula-
shunted via highly conductive skin tissue, also extends to tion, brain state, and task-induced state (e.g., Silvanto,
peripheral neuronal structures in the retina (Lorenz et al., Bona, Marelli, & Cattaneo, 2018; Schwarzkopf et al.,
2019; Schutter, 2016) and the vestibular system (Kwan, 2011). Such interactions complicate the conclusions
Forbes, Mitchell, Blouin, & Cullen, 2019). In addition to drawn from a given NIBS protocol. Moreover, learning
the E-field itself, physical side effects of the stimulation, or generalization effects influence performance, especially
such as the “click” sound and mechanical vibration gen- if the same task is measured repeatedly under different
erated by the discharging TMS coil, also affect mechano- stimulation conditions, as usually done for within-
receptors in the skin and reach the inner ear via both participant designs. Potential influences of these effects
airborne sound waves and bone conduction. Activation need to be controlled by counterbalancing the order of
of peripheral sensory structures then causes unintended tasks and stimulation conditions across participants. In
activation of primary and secondary sensory brain regions some learning paradigms, mirrored or alternative se-
and, eventually, also higher-order areas. Thus, both path- quences or stimulus lists can be used for a second ses-
ways (blue and green) can eventually activate remote re- sion. In case of implicit tasks or learning paradigms
gions and affect nontarget cognitive functions, for that cannot be repeated, between-participant designs
example, via the modulation of attentional orienting or need to be employed. Other unspecific effects such as
distraction of working memory content (Duecker & time of day and hormonal cycle may also influence

16 Journal of Cognitive Neuroscience Volume X, Number Y


cognitive function. Some of these factors can (and open but not eyes closed and thus during different alpha
should) be kept constant across experimental conditions, oscillation amplitudes (Neuling, Rach, & Herrmann,
for instance, by performing repeated measures in the 2013). In addition, classical offline NIBS protocols show
same participant at the same time of day. large intraindividual and interindividual variability in their
aftereffects, depending on several factors that vary across
sessions within an individual (e.g., the current brain state,
The Current Brain State
history of synaptic activity, hormonal levels, circadian
The magnitude and direction of NIBS effects may also rhythms) or across individuals (e.g., sex, age, individual
vary because of differences in the current brain state at depth and orientation of target region, genetics;
the time of stimulation, both within and between partic- Ridding & Ziemann, 2010), sometimes leading to null re-
ipants (Miniussi et al., 2013). The concept of “state de- sults (Beaulieu, Flamand, Massé-Alarie, & Schneider,
pendency” has been first introduced in the visual 2017; Heidegger et al., 2017). The individual variability
system (Silvanto, Muggleton, & Walsh, 2008; Silvanto, in response to offline TMS protocols in the motor system
Muggleton, Cowey, & Walsh, 2007), and state-dependent may be influenced by the specific interneuron networks
effects have been described across a variety of (cognitive) recruited, and corticospinal excitability measures may help
domains (Silvanto & Cattaneo, 2017). An impressive ex- to dissociate responders and nonresponders (Hamada,
ample for state dependency is the quantification of corti- Murase, Hasan, Balaratnam, & Rothwell, 2013). However,
cospinal excitability via the MEP acquired at rest and it is unclear how this translates to cognitive functions,
under precontraction. When compared to a relaxed tar- because excitability of M1 is not related to the responsive-
get muscle, slight precontraction leads to a considerable ness of other areas, and interindividual differences in
increase in the MEP size (Siebner, Hartwigsen, et al., cognitive abilities and functional organization further
2009). State dependency has also been demonstrated in contribute to large interindividual variability in response
remote network nodes by simultaneous TMS–fMRI when to NIBS protocols when targeting cognitive functions.
comparing the effects of different TMS intensities on neu-
ral activity at rest and during a grip task (e.g., Bestmann
et al., 2008). The key assumption is that the brain state
Metaplasticity
affects the distribution of excitability in the stimulated
population of neurons, which in turn affects their respon- Offline stimulation protocols may further be affected by
siveness to stimulation. A modulation of MEP amplitude metaplasticity, such that synaptic plasticity itself may vary
has also been shown as a function of the current ampli- depending on the history of a neuron’s postsynaptic ac-
tude and phase of neuronal oscillations during sleep tivity (Abraham & Bear, 1996). Metaplasticity contributes
(Bergmann et al., 2012) and wakefulness (Bergmann to network function and behavior and may be homeostatic
et al., 2019; Thies, Zrenner, Ziemann, & Bergmann, or nonhomeostatic. Homeostatic metaplasticity has
2018). Such oscillatory modulations of cortical excitability been demonstrated in human M1 for several combina-
led to the idea of brain-state-dependent brain stimula- tions of plasticity-inducing NIBS protocols (Karabanov
tion, allowing to confine stimulation to a certain target et al., 2015; Müller-Dahlhaus & Ziemann, 2015). For
state (Bergmann, 2018). Brain state dynamics may also instance, application of the same priming and test theta-
change the NIBS interference effect, such that intensities, burst protocols may reverse the effect of the test pro-
which normally impair perception, suddenly have a facil- tocol, whereas priming with the opposite protocol
itatory effect (Silvanto et al., 2018; Silvanto & Cattaneo, was shown to increase the effect of a test protocol
2017). For instance, when applied immediately before (Mastroeni et al., 2013; Murakami, Müller-Dahlhaus, Lu,
or early during a task, TMS resulted in a priming effect & Ziemann, 2012). Notably, effects of metaplasticity in-
by increasing activity in the target area to a level optimal duced by two consecutive NIBS protocols strongly de-
for task performance (e.g., Klaus & Hartwigsen, 2019; pend (among other factors like stimulation intensity)
Töpper, Mottaghy, Brügmann, Noth, & Huber, 1998). on the timing between priming and test protocol
In contrast, TMS impaired performance when given dur- (Müller-Dahlhaus & Ziemann, 2015), as the subsequent
ing the same task ( Wassermann et al., 1999; Flitman combination of two similar theta-burst protocols may also
et al., 1998). The current brain state may also interact result in a nonhomeostatic additive effect on corticosp-
with the polarity in TDCS studies, which is particularly inal excitability (e.g., Goldsworthy, Müller-Dahlhaus,
crucial for learning studies that engage different training Ridding, & Ziemann, 2015). At the behavioral level, it
phases (Dockery, Hueckel-Weng, Birbaumer, & Plewnia, was shown that the capacity of the motor cortex to un-
2009). Although initially introduced for online TMS, state dergo LTP-like plasticity in response to paired associative
dependency has also been suggested to influence offline stimulation was abolished immediately after motor train-
NIBS protocols (Nguyen, Deng, & Reinhart, 2018; ing (Stefan et al., 2006). These findings are particularly
Miniussi et al., 2013). The facilitatory aftereffect of relevant for cognitive neuroscience studies because many
10-Hz TACS on subsequent EEG alpha band power, for training interventions are combined with NIBS protocols
example, is only evident when TACS is applied with eyes to augment the effect of the behavioral intervention, and

Bergmann and Hartwigsen 17


the principle of homeostatic metaplasticity may general- intertrial intervals and/or TMS trains may be adopted or
ize across cortical areas (e.g., Bocci et al., 2014; Gatica trials of similar conditions may be grouped in short
Tossi, Stude, Schwenkreis, Tegenthoff, & Dinse, 2013). blocks.
Yet, the optimal timing between task and stimulation re-
mains unknown for most interventions, and potential ef-
Sham TMS
fects of metaplasticity are usually ignored. The above-
discussed results suggest that metaplasticity may aug- Sham TMS is meant to mimic sensory costimulation and
ment, diminish, or even reverse the expected direction serve as a placebo condition to control for participants’
of a given NIBS protocol. Although it remains largely un- beliefs and expectations. Different sham TMS approaches
clear how this affects cognitive processes, such effects have been adopted, including physical separation of the
may contribute to unexpected results or null findings of coil from the scalp with a spacer, placing an additional
NIBS studies in the field of cognitive neuroscience. coil with 90° tilt on top of the active coil and selectively
discharging the former as a sham condition, slightly tilt-
ing a coil on the scalp to avoid stimulation of the under-
CONTROL CONDITIONS—HOW TO DESIGN A lying brain region, employing a commercially available
VALID NIBS STUDY sham coil, or developing other “realistic” sham condi-
tions. Moving the coil away from the scalp preserves
Having outlined the numerous factors potentially inter-
the airborne sound induced by coil charging but induces
rupting or confounding the chain of causation for a typ-
little or no bone conduction and completely lacks so-
ical cognitive neuroscience NIBS study, we now discuss
matosensory costimulation (ter Braack, de Vos, & van
the most important experimental control conditions.
Putten, 2015). Likewise, an additional coil on top or tilt-
Control conditions are used to control influences that
ing the coil such that only the edge touches the head
cannot be completely eliminated in NIBS experiments,
produces roughly comparable auditory inputs compared
such as the unintentional costimulation of nontarget
to active TMS, yet almost no somatosensory inputs be-
brain regions or networks and peripheral sensory struc-
cause of a lack of peripheral nerve stimulation. The same
tures, the influence of the participants’ expectations
holds true for conventional sham coils that reduce the
and beliefs regarding the NIBS procedure, and the inev-
effective stimulation intensity by shielding or opposed
itable dependence of task performance on additional
current flow, which may prevent effective stimulation of
nontarget cognitive functions. Importantly, the control
the target area but also of peripheral structures.
conditions included in an experiment determine the con-
Consequently, participant blinding is hard to achieve with
clusions that can be drawn from the results. The typical
these approaches. This is particularly problematic for
rationale of NIBS experiments is that, if a given NIBS pro-
within-participant designs often used in cognitive neuro-
tocol, applied at a certain time point relative to a task,
science studies. To overcome these limitations, “realistic”
affects behavioral performance, this is proof of the causal
sham conditions aim to also mimic the somatosensory
relevance of the targeted area and/or neuronal activity
side effects. State-of-the-art realistic sham conditions thus
pattern for the cognitive function operationalized by
combine individually adjusted auditory noise masking to
the task. Depending on the hypothesis, several explicit
attenuate the click sound for both active and sham con-
(or implicit) assumptions are made regarding the speci-
ditions (Duecker & Sack, 2015), foam padding beneath
ficity of the investigated brain–behavior relationship with
the coil to attenuate vibration, a TMS coil discharging at
respect to the targeted anatomical location, temporal
an ineffective distance to control for residual auditory in-
window, and/or oscillatory frequency, as well as the af-
put, and individually adjusted electric stimulation of the
fected aspects of the task. For feasibility reasons, it is
skin via surface electrodes beneath the TMS coil to mimic
rarely possible to realize all control conditions in a single
peripheral nerve stimulation (Conde et al., 2019). Yet, even
study, particularly not in a fully crossed factorial design.
if electric stimulation intensity is individually adjusted,
However, any specificity claim needs to be explicitly
experienced participants likely notice the difference
tested with an appropriate control condition.
between realistic sham and effective stimulation, as the
skin sensation of the electric stimulation will be different
(e.g., Mennemeier et al., 2009; Rossi et al., 2007). This is
Stimulation-free Condition
particularly problematic when online TMS is applied at
Earlier studies often included trials without stimulation higher intensities to areas where the stimulation may
(i.e., “no TMS”) as a control condition, sometimes ran- yield unpleasant side effects (such as cranial/facial muscle
domly interleaved with effective stimulation to control twitches). These side effects can substantially confound
for carry-over and practice effects (Sandrini et al., the obtained results, as shown by a negative association
2011). However, this does not control for any unspecific between working memory performance and individual
side effects, and the difference between TMS and no TMS ratings of the unpleasantness of TMS (Abler et al.,
will be obvious for the participant. To avoid carry-over ef- 2005). An active control site is thus always preferable.
fects between trials or conditions, the duration of However, the choice of an active control site can be

18 Journal of Cognitive Neuroscience Volume X, Number Y


tricky, especially if complex cognitive functions are stud- respect to the target gyrus) can be sufficient to abolish
ied that are widely distributed across the brain, preclud- a specific effect, not only for MEPs evoked in M1 (Mills
ing the use of most well-matched regions as control sites. et al., 1992) but possibly also for other brain regions
(Thielscher et al., 2011; Thut, Veniero, et al., 2011).
Except for a simple dose reduction, inevitably changing
Sham TCS
the amount of sensory costimulation and, possibly, the
In contrast to most TMS studies, TCS studies usually in- participants’ beliefs regarding the effectiveness of the
clude sham conditions but no active control sites (i.e., stimulation, these controls can be considered high-level
control montages). Sham TCS is commonly realized by control conditions matching the effective experimental
ramping up the current to target intensity for, for exam- protocol relatively well, although they come with the risk
ple, 10–30 sec and immediately ramping it down again of residual transcranial effects on the brain.
without any further active stimulation, thus producing
some cutaneous (tingling, itching, burning) sensations
Control Tasks (Task Specificity)
in the beginning, when they are also strongest for real
TDCS before they habituate, supposedly making real Task specificity implies that, if a Cortical Area X is rele-
and sham TDCS indistinguishable (Gandiga, Hummel, vant for a Cognitive Process A but not B, then a given
& Cohen, 2006). However, recent evidence shows that, NIBS protocol over Area X should selectively modulate
even if low stimulation intensities around 1 mA are used, Task A but not B (Miniussi et al., 2013). This requires a
participant blinding is compromised (Greinacher, Buhôt, control task that differs from the task of interest selec-
Möller, & Learmonth, 2019; Turi et al., 2019). A recently tively in the cognitive function of interest but is matched
introduced sham TCS protocol combines a multielec- with respect to task difficulty, low-level sensory input,
trode montage with controlled shunting of currents via and supporting cognitive functions (e.g., perceptual, at-
a model-based quantification of transcutaneous and tentional, working memory, executive, or motor de-
transcranial effects, ensuring constant scalp sensations mands). Without evidence for task specificity, a NIBS
across the whole stimulation procedure and similar sen- protocol may simply interfere with any of those cognitive
sations relative to effective TCS (Neri et al., 2020). This functions instead, potentially resulting in the same effect
protocol was suggested to be superior in participant on behavioral performance. Importantly, some tasks
blinding relative to conventional bifocal ramp-up, ramp- (e.g., visual attention or orienting tasks) may also be
down sham protocols and may provide a realistic sham more strongly affected by sensory costimulation.
condition for TCS. As for TMS, carefully matched active
control montages are the gold standard also for TCS stud-
Control Regions (Anatomical Specificity)
ies, and an additional sham session with the same mon-
tage should serve only as low-level baseline. In TDCS Anatomical specificity means that a NIBS protocol only
studies, polarity specificity can be tested additionally by affects Task A when applied to Area X but not Area Y, be-
applying both anodal and cathodal TDCS with respect cause only the former causally contributes to the cogni-
to the target area of interest. tive function of interest. An active TMS control site or
TCS electrode montage is thus needed to demonstrate
that the observed effects actually depend on the stimula-
Control via Ineffective Stimulation Protocols
tion of a specific brain region and not only of the brain
For some NIBS protocols, it is possible to create an inef- per se or its sensory input structures. Numerous studies
fective stimulation protocol, which can be applied to the have used the vertex as an active control site, with
same region without causing the crucial neuronal effects. the rationale that auditory and somatosensory inputs
For example, a change in intertrain duration made the in- should be roughly similar to that of other target sites,
termediate theta burst stimulation protocol ineffective but the brain tissue is located deeper under the scalp
(Huang et al., 2005), and replacing theta bursts (triplets) and thought to mainly contain sensorimotor representa-
with single pulses also produced no aftereffects ( Volman, tions of the lower body, thus not influencing cognitive task
Roelofs, Koch, Verhagen, & Toni, 2011), since 200 pulses performance ( Jung, Bungert, Bowtell, & Jackson, 2016;
at 5 Hz have likely no lasting effect (Peinemann et al., Duecker, de Graaf, Jacobs, & Sack, 2013). However, de-
2004). The paired associative stimulation protocol pending on the target area of interest, differences in lat-
(Stefan, Kunesch, Cohen, Benecke, & Classen, 2000) be- eralization and unpleasantness are a potential issue. For
comes ineffective when using, imperceptibly for the par- instance, lateralized sham TMS was shown to pull covert
ticipant, a random selection of per se ineffective ISIs spatial attention toward the corresponding side of space,
(Bergmann et al., 2008). In general, a dose reduction thereby facilitating target detection in this hemifield
can make an ineffective protocol, for example, by reduc- (Duecker & Sack, 2013), which can hardly be achieved
ing stimulation intensity, number of pulses, or total dura- with vertex stimulation. TMS of pFC and (anterior) tem-
tion of the application. In addition, a change in TMS coil poral cortex can be particularly unpleasant because of
orientation (usually from orthogonal to parallel with costimulation of the facial nerves and muscles, and vertex

Bergmann and Hartwigsen 19


stimulation might not be an adequate control site for these et al., 1997). If TMS occurs after the stimulus in some
areas. Moreover, using simultaneous TMS–fMRI, it was dem- trials, response speed can be delayed because of expec-
onstrated that suprathreshold low-frequency rTMS ap- tancy violations causing the participant to “wait” for the
plied over the vertex induced widespread deactivations in TMS pulse (de Graaf, Jacobs, Roebroeck, & Sack, 2009).
the default mode network, although the BOLD signal in Again, temporal controls cannot replace an active control
the stimulated area was not affected ( Jung et al., 2016). site, as they are designed to reveal different specificities.
Although the origin of these remote effects remains un-
clear, the implication is that the vertex may not be a suit-
Control Frequencies (Frequency Specificity)
able control site for tasks that involve or interact with the
default mode network. Given the complex interaction of Frequency specificity is ignored in most NIBS studies.
the task positive network and the default mode network, However, for some research questions, this may be cru-
this would preclude the use of vertex stimulation for cial, especially if conclusions are drawn about entrain-
most cognitive functions, especially those directly associ- ment or plastic aftereffects of rhythmic NIBS protocols.
ated with default mode engagement such as semantic pro- For instance, in a hypothetical experiment, the conclusion
cessing, social cognition, autobiographical memory, self- that beta-TACS over the pFC affects working memory
related thinking, and consciousness (e.g., Binder, Desai, would only be valid if one could show these effects to
Graves, & Conant, 2009; Buckner, Andrews-Hanna, & be frequency specific (including control frequencies), in
Schacter, 2008). For some areas, homologous regions addition to the anatomical specificity (including a control
may be adequate active control sites, especially if lateral- montage and sham stimulation as baseline) and task
ization of a particular cognitive function is of interest. Yet, specificity (including a well-matched control task).
TMS may also affect contralateral areas via transcallosal Indeed, frequency specificity is often considered in cog-
connections (cf. Arrow 3), especially if high intensities or nitive neuroscience studies employing TACS, but many
frequencies are used. For some cognitive processes, later- studies still simply compare a frequency of interest
alization is less clear, and the homologous region might against sham stimulation, which does not justify any
contribute to a given task. The choice of an active control conclusion regarding the relevance of the stimulation
site that is matched for the most important dimensions can frequency. Some studies have explicitly employed
be guided by a recently introduced atlas for TMS studies symmetrical control frequencies below and above the tar-
(Meteyard & Holmes, 2018). Also for TCS, control get frequency (Herring et al., 2019; Romei et al., 2016) to
montages, well matched for peripheral costimulation prevent confounds with stimulation duration or number
effects, should be mandatory, allowing at least some of cycles. Yet, no clear procedure has been established to
degree of anatomical specificity to be claimed. For offline define the number of control frequencies or the distance
NIBS protocols, different stimulation sites should be from the frequency of interest (Herrmann et al., 2013).
targeted in different sessions several days apart to Obviously, the control frequencies should not be en-
prevent any carry-over effects of the stimulation. gaged in the task of interest. Individual adjustment of
the stimulation frequency may be favorable, at least in
the alpha band, although it is still unclear whether stim-
Control Time Points (Temporal Specificity)
ulation is more effective when it matches the eigenfre-
Depending on the research question and the applied quency of the brain (Reato, Rahman, Bikson, & Parra,
NIBS protocol, temporal specificity can be crucial. 2013) or is slightly different ( Vossen et al., 2015;
Indeed, chronometric approaches have substantially ad- Helfrich, Knepper, et al., 2014). Notably, TACS can in-
vanced the current knowledge on the time course of dif- duce rhythmic retinal phosphenes and cutaneous sensa-
ferent cognitive processes such as visual perception tions at most frequencies, although with different
(Amassian et al., 1989), visual orientation and awareness thresholds (Kanai, Chaieb, Antal, Walsh, & Paulus,
(de Graaf, Duecker, Fernholz, & Sack, 2015; Jacobs, 2008), highlighting again that frequency controls alone
Goebel, & Sack, 2012), motion-driven attention are not sufficient and control montages need to be
(Alexander, Laycock, Crewther, & Crewther, 2018), employed.
sound localization (At, Spierer, & Clarke, 2011), working
memory (Mottaghy, Gangitano, Krause, & Pascual-Leone,
Further Considerations
2003), or language (Schuhmann, Schiller, Goebel, &
Sack, 2012). Most chronometric TMS studies that system- To guarantee accurate coil or electrode placement and
atically vary the time point of stimulation argue that no maintenance across the experiment and avoid confounds
control site is needed because specificity is shown by induced by movement of the stimulation device, the use
the difference between time points. However, this as- of a stereotactic neuronavigation system and individual
sumption is invalid, because online TMS given immedi- T1-weighted images is highly recommended. In general,
ately before or with stimulus onset acts as an alerting TCS studies benefit from optimized montages or compu-
signal and causes unspecific intersensory facilitation ef- tationally optimized multichannel arrangements that may
fects on response speed (Duecker et al., 2013; Terao help to focalize the stimulated area and minimize

20 Journal of Cognitive Neuroscience Volume X, Number Y


unwanted peripheral costimulation (e.g., Khatoun et al., target region/network and that no unintended coac-
2018). E-field modeling should be employed to estimate tivations of other regions/networks can confound
the focality of electrode montages or coil placement, the results. This also allows to relate NIBS-induced
minimize the impact of noncortical stimulation, and op- neuronal and behavioral effects.
timize stimulation efficiency. As many other studies in 6. Include a control site! An active control site (TMS) or
the field of cognitive neuroscience, NIBS studies often montage (TCS), well matched for sensations and an-
suffer from relatively small and homogeneous samples noyance, is a strong control for sensory costimula-
(i.e., mainly healthy young student volunteers, with sam- tion confounds and serves to establish anatomical
ple sizes < 30 participants), which are not representative specificity. An additional (realistic) sham condition
of the general population. A priori calculations of statisti- as baseline is optimal.
cal power may guide sample size selection, and we prin- 7. Include a control task! An appropriate control task,
cipally advocate larger sample sizes, but for most NIBS which does not involve the cognitive function of in-
studies, the effect size is unknown beforehand. Whenever terest but is matched for task difficulty, is a strong
possible, NIBS experiments should be conducted as a control for possible confounding via coaffected sup-
within-participant design to reduce interindividual vari- porting cognitive functions (e.g., attention) and
ability and confounds based on imperfect randomization serves to establish task specificity.
of group membership. If control conditions are realized 8. Include control time windows (if applicable)! If the
in a between-participant design (e.g., for a post hoc control timing of the target neuronal process matters (e.g.,
experiment) the control condition needs to have the same when doing mental chronometry), stimulation at
sample size as the experimental condition to rule out different time points/windows is necessary to estab-
statistical power as a confound. lish temporal specificity.
9. Include control frequencies (if applicable)! If you
want to demonstrate entrainment or similar
Twelve General Recommendations for Designing frequency-specific effects, neighboring control fre-
Valid NIBS Studies quencies (ideally both lower and higher) are manda-
tory. An arrhythmic control stimulation can control
As a summary, we make the following 12 general recom-
for the number of stimulation pulses/cycles and the
mendations to be considered when designing a NIBS
sheer presence of rhythmicity but cannot establish
study in cognitive neuroscience (the applicability of these
frequency specificity.
recommendations may vary depending on the specifics
10. Reduce variability wherever possible! Within-
of your research question as well as the technical/logistic
participant designs reduce interindividual variability,
feasibility in your laboratory):
but between-participant designs are necessary
1. Know your target! Ensure you have identified the when repeated stimulation or task performance is
stimulation target with a spatial precision appropri- problematic for reasons of blinding or learning ef-
ate for your NIBS technique of choice. Consider an fects. Internal and external contextual factors (e.g.,
fMRI-based localizer for target identification if time of day, arousal) should be kept as comparable
possible. as possible between conditions.
2. Simulate the E-field! Use individual E-field modeling 11. Prevent carry-over and order effects! For offline
based on realistic MR-based head models to corrob- NIBS protocols, control conditions need to be con-
orate that the E-field is maximal at the target site ducted in separate sessions several days apart and
(sensitivity) and as limited to it as possible counterbalanced to ensure that previously induced
(specificity). synaptic plasticity cannot systematically interact
3. Adjust stimulation intensity! Stimulation intensity with the current protocol.
should be adjusted individually, taking coil–cortex 12. Choose your DV wisely! Thoroughly consider the
distance into account, even if no ideal reference val- outcome measure best reflecting the expected
ue yet exists (% RMT is established). The estimation change in cognitive function. Titrate task difficulty
of induced E-field strength is not (yet) state-of-the- individually to a level where performance becomes
art but recommended. sensitive to even small disturbances of underlying
4. Use neuronavigation! Individual MR-based neurona- computation.
vigation ensures precise TMS coil/TCS electrode
placement and maintenance within and across ses-
sions and is a basic prerequisite for high-quality
Conclusion
TMS and hd-TCS studies.
5. Combine NIBS with neuroimaging! Use neuroimag- We have outlined major challenges and potential pitfalls
ing (e.g., fMRI, EEG/MEG) to establish proof or tar- for experimentally testing and interpreting the chain of
get engagement, that is, demonstrate that the causation for NIBS studies in cognitive neuroscience.
desired neuronal effects have been induced in the We hope to have raised awareness for the potential

Bergmann and Hartwigsen 21


confounds and provided a guide for designing valid NIBS stimulation in mouse, monkey, and human. Neuroimage,
experiments. On the basis of the above-discussed stud- 194, 136–148.
Alekseichuk, I., Turi, Z., Amador de Lara, G., Antal, A., & Paulus,
ies, a promising avenue for the future will be the multi- W. (2016). Spatial working memory in humans depends on
method combination of NIBS with computational theta and high gamma synchronization in the prefrontal
modeling and neuroimaging to map stimulation-induced cortex. Current Biology, 26, 1513–1521.
changes at the neuronal and network level and link these Alexander, B., Laycock, R., Crewther, D. P., & Crewther, S. G.
changes with cognitive and behavioral effects. With re- (2018). An fMRI-neuronavigated chronometric TMS
investigation of V5 and intraparietal cortex in motion driven
spect to computational modeling approaches, particular attention. Frontiers in Human Neuroscience, 11, 638.
challenges include the modulation of the dynamics of Amassian, V. E., Cracco, R. Q., Maccabee, P. J., Cracco, J. B.,
long-term plastic aftereffects of different NIBS protocols Rudell, A., & Eberle, L. (1989). Suppression of visual
and the transfer from relatively easy decision-making pro- perception by magnetic coil stimulation of human occipital
cesses to more complex cognitive functions (e.g., lan- cortex. Electroencephalography and Clinical
Neurophysiology, 74, 458–462.
guage, social cognition, or problem solving). Recent Amassian, V. E., Cracco, R. Q., Maccabee, P. J., Cracco, J. B.,
advances in behavioral modeling may further help to Rudell, A. P., & Eberle, L. (1998). Transcranial magnetic
bridge the gap between cognitive theories and behavioral stimulation in study of the visual pathway. Journal of Clinical
outcome measures of NIBS experiments, and the inclu- Neurophysiology, 15, 288–304.
sion of more natural tasks or stimuli will increase the eco- Antal, A., Boros, K., Poreisz, C., Chaieb, L., Terney, D., & Paulus,
W. (2008). Comparatively weak after-effects of transcranial
logical validity. To deepen our understanding of the alternating current stimulation (tACS) on cortical excitability
modulatory effects of NIBS protocols, future studies in humans. Brain Stimulation, 1, 97–105.
should relate the simulated strength of the induced E- Antal, A., & Herrmann, C. S. (2016). Transcranial alternating
field in the target area to neuroimaging-based assess- current and random noise stimulation: Possible mechanisms.
ments of target engagement and behavioral outcome Neural Plasticity, 2016, 3616807.
Antal, A., Polania, R., Schmidt-Samoa, C., Dechent, P., & Paulus,
measures. As NIBS studies are complex and time con- W. (2011). Transcranial direct current stimulation over the
suming and the modulatory effects are often small and primary motor cortex during fMRI. Neuroimage, 55,
variable across participants, a particular challenge for 590–596.
the future is the inclusion of larger sample sizes to guar- Asamoah, B., Khatoun, A., & McLaughlin, M. (2019). tACS motor
antee sufficient experimental power. Here, multicenter system effects can be caused by transcutaneous stimulation
of peripheral nerves. Nature Communications, 10, 266.
approaches may identify valid paradigms for NIBS studies At, A., Spierer, L., & Clarke, S. (2011). The role of the right
that may provide further insight into causal structure– parietal cortex in sound localization: A chronometric single
function relationships. pulse transcranial magnetic stimulation study.
Neuropsychologia, 49, 2794–2797.
Barker, A. T., Jalinous, R., & Freeston, I. L. (1985). Non-invasive
Reprint requests should be sent to Til Ole Bergmann, Leibniz magnetic stimulation of human motor cortex. Lancet, 325,
Institute for Resilience Research, Wallstraße 7-9, 55122 Mainz, 1106–1107.
Germany, or via e-mail: til-ole.bergmann@lir-mainz.de. Batsikadze, G., Moliadze, V., Paulus, W., Kuo, M.-F., & Nitsche,
M. A. (2013). Partially non-linear stimulation intensity-
dependent effects of direct current stimulation on motor
cortex excitability in humans. Journal of Physiology, 591,
REFERENCES 1987–2000.
Beaulieu, L.-D., Flamand, V. H., Massé-Alarie, H., & Schneider,
Aberra, A. S., Wang, B., Grill, W. M., & Peterchev, A. V. (2020). C. (2017). Reliability and minimal detectable change of
Simulation of transcranial magnetic stimulation in head transcranial magnetic stimulation outcomes in healthy
model with morphologically-realistic cortical neurons. Brain adults: A systematic review. Brain Stimulation, 10,
Stimulation, 13, 175–189. 196–213.
Abler, B., Walter, H., Wunderlich, A., Grothe, J., Schönfeldt- Bergmann, T. O. (2018). Brain state-dependent brain
Lecuona, C., Spitzer, M., et al. (2005). Side effects of stimulation. Frontiers in Psychology, 9, 2108.
transcranial magnetic stimulation biased task performance in Bergmann, T. O., & Born, J. (2018). Phase–amplitude coupling:
a cognitive neuroscience study. Brain Topography, 17, A general mechanism for memory processing and synaptic
193–196. plasticity? Neuron, 97, 10–13.
Abraham, W. C., & Bear, M. F. (1996). Metaplasticity: The Bergmann, T. O., Karabanov, A., Hartwigsen, G., Thielscher, A.,
plasticity of synaptic plasticity. Trends in Neurosciences, 19, & Siebner, H. R. (2016). Combining non-invasive transcranial
126–130. brain stimulation with neuroimaging and electrophysiology:
Abrahamyan, A., Clifford, C. W. G., Arabzadeh, E., & Harris, J. A. Current approaches and future perspectives. Neuroimage,
(2011). Improving visual sensitivity with subthreshold 140, 4–19.
transcranial magnetic stimulation. Journal of Neuroscience, Bergmann, T. O., Lieb, A., Zrenner, C., & Ziemann, U. (2019).
31, 3290–3294. Pulsed facilitation of corticospinal excitability by the
Alekseichuk, I., Falchier, A. Y., Linn, G., Xu, T., Milham, M. P., sensorimotor μ-alpha rhythm. Journal of Neuroscience,
Schroeder, C. E., et al. (2019). Electric field dynamics in the 39, 10034–10043.
brain during multi-electrode transcranial electric stimulation. Bergmann, T. O., Mölle, M., Marshall, L., Kaya-Yildiz, L., Born, J.,
Nature Communications, 10, 2573. & Siebner, H. R. (2008). A local signature of LTP- and LTD-
Alekseichuk, I., Mantell, K., Shirinpour, S., & Opitz, A. (2019). like plasticity in human NREM sleep. European Journal of
Comparative modeling of transcranial magnetic and electric Neuroscience, 27, 2241–2249.

22 Journal of Cognitive Neuroscience Volume X, Number Y


Bergmann, T. O., Mölle, M., Schmidt, M. A., Lindner, C., Chambliss, B. (2018). The mind–body problem. Wiley
Marshall, L., Born, J., et al. (2012). EEG-guided transcranial Interdisciplinary Reviews: Cognitive Science, 9, e1463.
magnetic stimulation reveals rapid shifts in motor cortical Chen, J.-Y., Chauvette, S., Skorheim, S., Timofeev, I., &
excitability during the human sleep slow oscillation. Journal Bazhenov, M. (2012). Interneuron-mediated inhibition
of Neuroscience, 32, 243–253. synchronizes neuronal activity during slow oscillation.
Bestmann, S., Baudewig, J., Siebner, H. R., Rothwell, J. C., & Journal of Physiology, 590, 3987–4010.
Frahm, J. (2003). Subthreshold high-frequency TMS of Chen, R., Lozano, A. M., & Ashby, P. (1999). Mechanism of the
human primary motor cortex modulates interconnected silent period following transcranial magnetic stimulation:
frontal motor areas as detected by interleaved fMRI–TMS. Evidence from epidural recordings. Experimental Brain
Neuroimage, 20, 1685–1696. Research, 128, 539–542.
Bestmann, S., Baudewig, J., Siebner, H. R., Rothwell, J. C., & Cohen, M. X. (2017). Where does EEG come from and what
Frahm, J. (2005). BOLD MRI responses to repetitive TMS over does it mean? Trends in Neurosciences, 40, 208–218.
human dorsal premotor cortex. Neuroimage, 28, 22–29. Conde, V., Tomasevic, L., Akopian, I., Stanek, K., Saturnino, G. B.,
Bestmann, S., & Feredoes, E. (2013). Combined Thielscher, A., et al. (2019). The non-transcranial TMS-
neurostimulation and neuroimaging in cognitive evoked potential is an inherent source of ambiguity in TMS–
neuroscience: Past, present, and future. Annals of the New EEG studies. Neuroimage, 185, 300–312.
York Academy of Sciences, 1296, 11–30. Datta, A., Bansal, V., Diaz, J., Patel, J., Reato, D., & Bikson, M.
Bestmann, S., Swayne, O., Blankenburg, F., Ruff, C. C., Haggard, (2009). Gyri-precise head model of transcranial direct current
P., Weiskopf, N., et al. (2008). Dorsal premotor cortex exerts stimulation: Improved spatial focality using a ring electrode
state-dependent causal influences on activity in contralateral versus conventional rectangular pad. Brain Stimulation, 2,
primary motor and dorsal premotor cortex. Cerebral Cortex, 201–207.
18, 1281–1291. Davare, M., Lemon, R., & Olivier, E. (2008). Selective
Binder, J. R., Desai, R. H., Graves, W. W., & Conant, L. L. (2009). modulation of interactions between ventral premotor cortex
Where is the semantic system? A critical review and meta- and primary motor cortex during precision grasping in
analysis of 120 functional neuroimaging studies. Cerebral humans. Journal of Physiology, 586, 2735–2742.
Cortex, 19, 2767–2796. de Graaf, T. A., Duecker, F., Fernholz, M. H. P., & Sack, A. T.
Bindman, L. J., Lippold, O. C. J., & Redfearn, J. W. T. (1964). The (2015). Spatially specific vs. unspecific disruption of visual
action of brief polarizing currents on the cerebral cortex of orientation perception using chronometric pre-stimulus
the rat (1) during current flow and (2) in the production of TMS. Frontiers in Behavioral Neuroscience, 9, 5.
long-lasting after-effects. Journal of Physiology, 172, de Graaf, T. A., Jacobs, C., Roebroeck, A., & Sack, A. T. (2009).
369–382. fMRI effective connectivity and TMS chronometry:
Bocci, T., Caleo, M., Tognazzi, S., Francini, N., Briscese, L., Complementary accounts of causality in the visuospatial
Maffei, L., et al. (2014). Evidence for metaplasticity in the judgment network. PLoS One, 4, e8307.
human visual cortex. Journal of Neural Transmission, 121, Dijkstra, N., & de Bruin, L. (2016). Cognitive neuroscience and
221–231. causal inference: Implications for psychiatry. Frontiers in
Bonaiuto, J. J., & Bestmann, S. (2015). Understanding the Psychiatry, 7, 129.
nonlinear physiological and behavioral effects of tDCS Di Lazzaro, V., & Ziemann, U. (2013). The contribution of
through computational neurostimulation. Progress in Brain transcranial magnetic stimulation in the functional evaluation
Research, 222, 75–103. of microcircuits in human motor cortex. Frontiers in Neural
Bonaiuto, J. J., de Berker, A., & Bestmann, S. (2016). Response Circuits, 7, 18.
repetition biases in human perceptual decisions are Di Lazzaro, V., Ziemann, U., & Lemon, R. N. (2008). State of the
explained by activity decay in competitive attractor models. art: Physiology of transcranial motor cortex stimulation.
eLife, 5, e20047. Brain Stimulation, 1, 345–362.
Brasil-Neto, J. P., McShane, L. M., Fuhr, P., Hallett, M., & Cohen, Dockery, C. A., Hueckel-Weng, R., Birbaumer, N., & Plewnia, C.
L. G. (1992). Topographic mapping of the human motor (2009). Enhancement of planning ability by transcranial direct
cortex with magnetic stimulation: Factors affecting accuracy current stimulation. Journal of Neuroscience, 29, 7271–7277.
and reproducibility. Electroencephalography and Clinical Duecker, F., de Graaf, T. A., Jacobs, C., & Sack, A. T. (2013).
Neurophysiology, 85, 9–16. Time- and task-dependent non-neural effects of real and
Buckner, R. L., Andrews-Hanna, J. R., & Schacter, D. L. (2008). sham TMS. PLoS One, 8, e73813.
The brain’s default network: Anatomy, function, and Duecker, F., Frost, M. A., de Graaf, T. A., Graewe, B., Jacobs, C.,
relevance to disease. Annals of the New York Academy of Goebel, R., et al. (2014). The cortex-based alignment
Sciences, 1124, 1–38. approach to TMS coil positioning. Journal of Cognitive
Bungert, A., Antunes, A., Espenhahn, S., & Thielscher, A. Neuroscience, 26, 2321–2329.
(2017). Where does TMS stimulate the motor cortex? Duecker, F., & Sack, A. T. (2013). Pre-stimulus sham TMS
Combining electrophysiological measurements and realistic facilitates target detection. PLoS One, 8, e57765.
field estimates to reveal the affected cortex position. Duecker, F., & Sack, A. T. (2015). Rethinking the role of sham
Cerebral Cortex, 27, 5083–5094. TMS. Frontiers in Psychology, 6, 210.
Buzsáki, G., & Draguhn, A. (2004). Neuronal oscillations in Edelman, G. M., & Gally, J. A. (2001). Degeneracy and
cortical networks. Science, 304, 1926–1929. complexity in biological systems. Proceedings of the
Capotosto, P., Babiloni, C., Romani, G. L., & Corbetta, M. National Academy of Sciences, U.S.A., 98, 13763–13768.
(2012). Differential contribution of right and left parietal Ellison, A., & Cowey, A. (2009). Differential and co-involvement
cortex to the control of spatial attention: A simultaneous of areas of the temporal and parietal streams in visual tasks.
EEG–rTMS study. Cerebral Cortex, 22, 446–454. Neuropsychologia, 47, 1609–1614.
Cattaneo, L., Barchiesi, G., Tabarelli, D., Arfeller, C., Sato, M., & Ermentrout, G. B., Galán, R. F., & Urban, N. N. (2008).
Glenberg, A. M. (2011). One’s motor performance Reliability, synchrony and noise. Trends in Neurosciences,
predictably modulates the understanding of others’ actions 31, 428–434.
through adaptation of premotor visuo-motor neurons. Social Esmaeilpour, Z., Marangolo, P., Hampstead, B. M., Bestmann,
Cognitive and Affective Neuroscience, 6, 301–310. S., Galletta, E., Knotkova, H., et al. (2018). Incomplete

Bergmann and Hartwigsen 23


evidence that increasing current intensity of tDCS boosts human motor cortical plasticity. Cerebral Cortex, 23,
outcomes. Brain Stimulation, 11, 310–321. 1593–1605.
Faisal, A. A., Selen, L. P. J., & Wolpert, D. M. (2008). Noise in the Hämmerer, D., Bonaiuto, J., Klein-Flügge, M., Bikson, M., &
nervous system. Nature Reviews Neuroscience, 9, 292–303. Bestmann, S. (2016). Selective alteration of human value
Ferbert, A., Priori, A., Rothwell, J. C., Day, B. L., Colebatch, J. G., decisions with medial frontal tDCS is predicted by changes
& Marsden, C. D. (1992). Interhemispheric inhibition of the in attractor dynamics. Scientific Reports, 6, 25160.
human motor cortex. Journal of Physiology, 453, 525–546. Hanslmayr, S., Matuschek, J., & Fellner, M.-C. (2014).
Fertonani, A., & Miniussi, C. (2017). Transcranial electrical Entrainment of prefrontal beta oscillations induces an
stimulation: What we know and do not know about endogenous echo and impairs memory formation. Current
mechanisms. Neuroscientist, 23, 109–123. Biology, 24, 904–909.
Filmer, H. L., Mattingley, J. B., & Dux, P. E. (2020). Modulating Hanslmayr, S., Staudigl, T., & Fellner, M.-C. (2012). Oscillatory
brain activity and behaviour with tDCS: Rumours of its death power decreases and long-term memory: The information via
have been greatly exaggerated. Cortex, 123, 141–151. desynchronization hypothesis. Frontiers in Human
Fitzgerald, P. B., Fountain, S., & Daskalakis, Z. J. (2006). A Neuroscience, 6, 74.
comprehensive review of the effects of rTMS on motor Harquel, S., Bacle, T., Beynel, L., Marendaz, C., Chauvin, A., &
cortical excitability and inhibition. Clinical Neurophysiology, David, O. (2016). Mapping dynamical properties of cortical
117, 2584–2596. microcircuits using robotized TMS and EEG: Towards
Flitman, S. S., Grafman, J., Wassermann, E. M., Cooper, V., functional cytoarchitectonics. Neuroimage, 135, 115–124.
O’Grady, J., Pascual-Leone, A., et al. (1998). Linguistic Hartwigsen, G. (2018). Flexible redistribution in cognitive
processing during repetitive transcranial magnetic networks. Trends in Cognitive Sciences, 22, 687–698.
stimulation. Neurology, 50, 175–181. Hartwigsen, G., Baumgaertner, A., Price, C. J., Koehnke, M.,
Fomenko, A., Neudorfer, C., Dallapiazza, R. F., Kalia, S. K., & Ulmer, S., & Siebner, H. R. (2010). Phonological decisions
Lozano, A. M. (2018). Low-intensity ultrasound require both the left and right supramarginal gyri.
neuromodulation: An overview of mechanisms and emerging Proceedings of the National Academy of Sciences, U.S.A.,
human applications. Brain Stimulation, 11, 1209–1217. 107, 16494–16499.
Fries, P. (2015). Rhythms for cognition: Communication Hartwigsen, G., Bergmann, T. O., Herz, D. M., Angstmann, S.,
through coherence. Neuron, 88, 220–235. Karabanov, A., Raffin, E., et al. (2015). Modeling the effects of
Fröhlich, F., & McCormick, D. A. (2010). Endogenous electric noninvasive transcranial brain stimulation at the biophysical,
fields may guide neocortical network activity. Neuron, 67, network, and cognitive level. Progress in Brain Research,
129–143. 222, 261–287.
Gandiga, P. C., Hummel, F. C., & Cohen, L. G. (2006). Hartwigsen, G., Bestmann, S., Ward, N. S., Woerbel, S.,
Transcranial DC stimulation (tDCS): A tool for double-blind Mastroeni, C., Granert, O., et al. (2012). Left dorsal premotor
sham-controlled clinical studies in brain stimulation. Clinical cortex and supramarginal gyrus complement each other
Neurophysiology, 117, 845–850. during rapid action reprogramming. Journal of
Gatica Tossi, M. A., Stude, P., Schwenkreis, P., Tegenthoff, M., & Neuroscience, 32, 16162–16171.
Dinse, H. R. (2013). Behavioural and neurophysiological Hartwigsen, G., Bzdok, D., Klein, M., Wawrzyniak, M., Stockert,
markers reveal differential sensitivity to homeostatic A., Wrede, K., et al. (2017). Rapid short-term reorganization
interactions between centrally and peripherally applied in the language network. eLife, 6, e25964.
passive stimulation. European Journal of Neuroscience, Hartwigsen, G., Price, C. J., Baumgaertner, A., Geiss, G.,
38, 2893–2901. Koehnke, M., Ulmer, S., et al. (2010). The right posterior
Gebodh, N., Esmaeilpour, Z., Adair, D., Chelette, K., inferior frontal gyrus contributes to phonological word
Dmochowski, J., Woods, A. J., et al. (2019). Inherent decisions in the healthy brain: Evidence from dual-site TMS.
physiological artifacts in EEG during tDCS. Neuroimage, Neuropsychologia, 48, 3155–3163.
185, 408–424. Hartwigsen, G., & Siebner, H. R. (2015). Joint contribution of
Goldsworthy, M. R., Müller-Dahlhaus, F., Ridding, M. C., & left dorsal premotor cortex and supramarginal gyrus to rapid
Ziemann, U. (2015). Resistant against de-depression: LTD- action reprogramming. Brain Stimulation, 8, 945–952.
like plasticity in the human motor cortex induced by spaced Haug, B. A., Schönle, P. W., Knobloch, C., & Köhne, M. (1992).
cTBS. Cerebral Cortex, 25, 1724–1734. Silent period measurement revives as a valuable diagnostic
Gomez-Tames, J., Hamasaka, A., Laakso, I., Hirata, A., & Ugawa, tool with transcranial magnetic stimulation.
Y. (2018). Atlas of optimal coil orientation and position for Electroencephalography and Clinical Neurophysiology, 85,
TMS: A computational study. Brain Stimulation, 11, 158–160.
839–848. Heidegger, T., Hansen-Goos, O., Batlaeva, O., Annak, O.,
Greinacher, R., Buhôt, L., Möller, L., & Learmonth, G. (2019). Ziemann, U., & Lötsch, J. (2017). A data-driven approach to
The time course of ineffective sham-blinding during low- responder subgroup identification after paired continuous
intensity (1 mA) transcranial direct current stimulation. theta burst stimulation. Frontiers in Human Neuroscience,
European Journal of Neuroscience, 50, 3380–3388. 11, 382.
Grosse-Wentrup, M., Janzing, D., Siegel, M., & Schölkopf, B. Helfrich, R. F., Knepper, H., Nolte, G., Strüber, D., Rach, S.,
(2016). Identification of causal relations in neuroimaging data Herrmann, C. S., et al. (2014). Selective modulation of
with latent confounders: An instrumental variable approach. interhemispheric functional connectivity by HD-tACS shapes
Neuroimage, 125, 825–833. perception. PLoS Biology, 12, e1002031.
Hallam, G. P., Whitney, C., Hymers, M., Gouws, A. D., & Helfrich, R. F., Schneider, T. R., Rach, S., Trautmann-Lengsfeld,
Jefferies, E. (2016). Charting the effects of TMS with fMRI: S. A., Engel, A. K., & Herrmann, C. S. (2014). Entrainment of
Modulation of cortical recruitment within the distributed brain oscillations by transcranial alternating current
network supporting semantic control. Neuropsychologia, stimulation. Current Biology, 24, 333–339.
93, 40–52. Herring, J. D., Esterer, S., Marshall, T. R., Jensen, O., &
Hamada, M., Murase, N., Hasan, A., Balaratnam, M., & Rothwell, Bergmann, T. O. (2019). Low-frequency alternating current
J. C. (2013). The role of interneuron networks in driving stimulation rhythmically suppresses gamma-band oscillations

24 Journal of Cognitive Neuroscience Volume X, Number Y


and impairs perceptual performance. Neuroimage, 184, language production. Human Brain Mapping, 40,
440–449. 3279–3287.
Herring, J. D., Thut, G., Jensen, O., & Bergmann, T. O. (2015). Kohl, C., Spieser, L., Forster, B., Bestmann, S., & Yarrow, K.
Attention modulates TMS-locked alpha oscillations in the (2019). The neurodynamic decision variable in human multi-
visual cortex. Journal of Neuroscience, 35, 14435–14447. alternative perceptual choice. Journal of Cognitive
Herrmann, C. S., Rach, S., Neuling, T., & Strüber, D. (2013). Neuroscience, 31, 262–277.
Transcranial alternating current stimulation: A review of the Krause, M. R., Vieira, P. G., Csorba, B. A., Pilly, P. K., & Pack, C. C.
underlying mechanisms and modulation of cognitive (2019). Transcranial alternating current stimulation
processes. Frontiers in Human Neuroscience, 7, 279. entrains single-neuron activity in the primate brain.
Hoogendam, J. M., Ramakers, G. M. J., & Di Lazzaro, V. (2010). Proceedings of the National Academy of Sciences, U.S.A.,
Physiology of repetitive transcranial magnetic stimulation of 116, 5747–5755.
the human brain. Brain Stimulation, 3, 95–118. Kwan, A., Forbes, P. A., Mitchell, D. E., Blouin, J.-S., & Cullen,
Huang, Y.-Z., Edwards, M. J., Rounis, E., Bhatia, K. P., & K. E. (2019). Neural substrates, dynamics and thresholds of
Rothwell, J. C. (2005). Theta burst stimulation of the human galvanic vestibular stimulation in the behaving primate.
motor cortex. Neuron, 45, 201–206. Nature Communications, 10, 1904.
Ilmoniemi, R. J., & Kičić, D. (2010). Methodology for combined Lepage, J.-F., Saint-Amour, D., & Théoret, H. (2008). EEG and
TMS and EEG. Brain Topography, 22, 233–248. neuronavigated single-pulse TMS in the study of the
Inghilleri, M., Berardelli, A., Cruccu, G., & Manfredi, M. (1993). observation/execution matching system: Are both techniques
Silent period evoked by transcranial stimulation of the measuring the same process? Journal of Neuroscience
human cortex and cervicomedullary junction. Journal of Methods, 175, 17–24.
Physiology, 466, 521–534. Li, B., Virtanen, J. P., Oeltermann, A., Schwarz, C., Giese, M. A.,
Jacobs, C., Goebel, R., & Sack, A. T. (2012). Visual awareness Ziemann, U., et al. (2017). Lifting the veil on the dynamics of
suppression by pre-stimulus brain stimulation; a neural neuronal activities evoked by transcranial magnetic
effect. Neuroimage, 59, 616–624. stimulation. eLife, 6, e30552.
Jensen, O., & Colgin, L. L. (2007). Cross-frequency coupling Lisman, J. E., & Jensen, O. (2013). The theta–gamma neural
between neuronal oscillations. Trends in Cognitive Sciences, code. Neuron, 77, 1002–1016.
11, 267–269. Liu, A., Vöröslakos, M., Kronberg, G., Henin, S., Krause, M. R.,
Jensen, O., Gips, B., Bergmann, T. O., & Bonnefond, M. (2014). Huang, Y., et al. (2018). Immediate neurophysiological
Temporal coding organized by coupled alpha and gamma effects of transcranial electrical stimulation. Nature
oscillations prioritize visual processing. Trends in Communications, 9, 5092.
Neurosciences, 37, 357–369. Lorenz, R., Simmons, L. E., Monti, R. P., Arthur, J. L., Limal, S.,
Jerusalem, A., Al-Rekabi, Z., Chen, H., Ercole, A., Malboubi, M., Laakso, I., et al. (2019). Efficiently searching through large
Tamayo-Elizalde, M., et al. (2019). Electrophysiological– tACS parameter spaces using closed-loop Bayesian
mechanical coupling in the neuronal membrane and its role optimization. Brain Stimulation, 12, 1484–1489.
in ultrasound neuromodulation and general anaesthesia. Luber, B., & Lisanby, S. H. (2014). Enhancement of human
Acta Biomaterialia, 97, 116–140. cognitive performance using transcranial magnetic
Jung, J., Bungert, A., Bowtell, R., & Jackson, S. R. (2016). Vertex stimulation (TMS). Neuroimage, 85, 961–970.
stimulation as a control site for transcranial magnetic Marshall, T. R., O’Shea, J., Jensen, O., & Bergmann, T. O.
stimulation: A concurrent TMS/fMRI study. Brain (2015). Frontal eye fields control attentional modulation of
Stimulation, 9, 58–64. alpha and gamma oscillations in contralateral occipitoparietal
Kaarre, O., Äikiä, M., Kallioniemi, E., Könönen, M., Kekkonen, cortex. Journal of Neuroscience, 35, 1638–1647.
V., Heikkinen, N., et al. (2018). Association of the N100 TMS- Massimini, M., Ferrarelli, F., Huber, R., Esser, S. K., Singh, H., &
evoked potential with attentional processes: A motor cortex Tononi, G. (2005). Breakdown of cortical effective
TMS–EEG study. Brain and Cognition, 122, 9–16. connectivity during sleep. Science, 309, 2228–2232.
Kammer, T. (1998). Phosphenes and transient scotomas Mastroeni, C., Bergmann, T. O., Rizzo, V., Ritter, C., Klein, C.,
induced by magnetic stimulation of the occipital lobe: Their Pohlmann, I., et al. (2013). Brain-derived neurotrophic factor
topographic relationship. Neuropsychologia, 37, 191–198. —A major player in stimulation-induced homeostatic
Kanai, R., Chaieb, L., Antal, A., Walsh, V., & Paulus, W. (2008). metaplasticity of human motor cortex? PLoS One, 8, e57957.
Frequency-dependent electrical stimulation of the visual McDonnell, M. N., Orekhov, Y., & Ziemann, U. (2006). The
cortex. Current Biology, 18, 1839–1843. role of GABAB receptors in intracortical inhibition in the
Kapur, N. (1996). Paradoxical functional facilitation in brain– human motor cortex. Experimental Brain Research, 173,
behaviour research: A critical review. Brain, 119, 1775–1790. 86–93.
Karabanov, A., Ziemann, U., Hamada, M., George, M. S., Mehler, D. M. A., & Kording, K. P. (2018). The lure of causal
Quartarone, A., Classen, J., et al. (2015). Consensus paper: statements: Rampant mis-inference of causality in estimated
Probing homeostatic plasticity of human cortex with non- connectivity. arXiv:1812.03363.
invasive transcranial brain stimulation. Brain Stimulation, 8, Mennemeier, M. S., Triggs, W. J., Chelette, K. C., Woods, A. J.,
442–454. Kimbrell, T. A., & Dornhoffer, J. L. (2009). Sham transcranial
Kasten, F. H., Duecker, K., Maack, M. C., Meiser, A., & magnetic stimulation using electrical stimulation of the scalp.
Herrmann, C. S. (2019). Integrating electric field modeling Brain Stimulation, 2, 168–173.
and neuroimaging to explain inter-individual variability of Meteyard, L., & Holmes, N. P. (2018). TMS SMART—Scalp
tACS effects. Nature Communications, 10, 5427. mapping of annoyance ratings and twitches caused by
Khatoun, A., Breukers, J., Op de Beeck, S., Nica, I. G., Aerts, transcranial magnetic stimulation. Journal of Neuroscience
J.-M., Seynaeve, L., et al. (2018). Using high-amplitude and Methods, 299, 34–44.
focused transcranial alternating current stimulation to entrain Mills, K. R., Boniface, S. J., & Schubert, M. (1992). Magnetic
physiological tremor. Scientific Reports, 8, 4927. brain stimulation with a double coil: The importance of coil
Klaus, J., & Hartwigsen, G. (2019). Dissociating semantic and orientation. Electroencephalography and Clinical
phonological contributions of the left inferior frontal gyrus to Neurophysiology, 85, 17–21.

Bergmann and Hartwigsen 25


Miniussi, C., Harris, J. A., & Ruzzoli, M. (2013). Modelling non- Opitz, A., Paulus, W., Will, S., Antunes, A., & Thielscher, A.
invasive brain stimulation in cognitive neuroscience. (2015). Determinants of the electric field during transcranial
Neuroscience & Biobehavioral Reviews, 37, 1702–1712. direct current stimulation. Neuroimage, 109, 140–150.
Miniussi, C., Ruzzoli, M., & Walsh, V. (2010). The mechanism of Opitz, A., Windhoff, M., Heidemann, R. M., Turner, R., &
transcranial magnetic stimulation in cognition. Cortex, 46, Thielscher, A. (2011). How the brain tissue shapes the
128–130. electric field induced by transcranial magnetic stimulation.
Moliadze, V., Atalay, D., Antal, A., & Paulus, W. (2012). Close to Neuroimage, 58, 849–859.
threshold transcranial electrical stimulation preferentially O’Shea, J., Johansen-Berg, H., Trief, D., Göbel, S., & Rushworth,
activates inhibitory networks before switching to excitation M. F. S. (2007). Functionally specific reorganization in human
with higher intensities. Brain Stimulation, 5, 505–511. premotor cortex. Neuron, 54, 479–490.
Moliadze, V., Zhao, Y., Eysel, U., & Funke, K. (2003). Effect of O’Shea, J., Revol, P., Cousijn, H., Near, J., Petitet, P., Jacquin-
transcranial magnetic stimulation on single-unit activity in the Courtois, S., et al. (2017). Induced sensorimotor cortex
cat primary visual cortex. Journal of Physiology, 553, plasticity remediates chronic treatment-resistant visual
665–679. neglect. eLife, 6, e26602.
Mottaghy, F. M., Gangitano, M., Krause, B. J., & Pascual-Leone, Parkin, B. L., Bhandari, M., Glen, J. C., & Walsh, V. (2019). The
A. (2003). Chronometry of parietal and prefrontal activations physiological effects of transcranial electrical stimulation do
in verbal working memory revealed by transcranial magnetic not apply to parameters commonly used in studies of
stimulation. Neuroimage, 18, 565–575. cognitive neuromodulation. Neuropsychologia, 128,
Möttönen, R., van de Ven, G. M., & Watkins, K. E. (2014). 332–339.
Attention fine-tunes auditory–motor processing of speech Pascual-Leone, A., Walsh, V., & Rothwell, J. (2000). Transcranial
sounds. Journal of Neuroscience, 34, 4064–4069. magnetic stimulation in cognitive neuroscience—Virtual
Mueller, J. K., Grigsby, E. M., Prevosto, V., Petraglia, F. W., III, lesion, chronometry, and functional connectivity. Current
Rao, H., Deng, Z.-D., et al. (2014). Simultaneous transcranial Opinion in Neurobiology, 10, 232–237.
magnetic stimulation and single-neuron recording in alert Payne, J. S., & Tainturier, M.-J. (2018). tDCS facilitation of
non-human primates. Nature Neuroscience, 17, 1130–1136. picture naming: Item-specific, task general, or neither?
Müller-Dahlhaus, F., & Ziemann, U. (2015). Metaplasticity in Frontiers in Neuroscience, 12, 549.
human cortex. Neuroscientist, 21, 185–202. Pearl, J. (2010). An introduction to causal inference.
Murakami, T., Müller-Dahlhaus, F., Lu, M.-K., & Ziemann, U. International Journal of Biostatistics, 6, 7.
(2012). Homeostatic metaplasticity of corticospinal excitatory Pearl, J., & Mackenzie, D. (2018). The book of why: The new
and intracortical inhibitory neural circuits in human motor science of cause and effect. New York: Basic Books.
cortex. Journal of Physiology, 590, 5765–5781. Peinemann, A., Reimer, B., Löer, C., Quartarone, A., Munchau, A.,
Murakami, T., Restle, J., & Ziemann, U. (2011). Observation– Conrad, B., et al. (2004). Long-lasting increase in
execution matching and action inhibition in human primary corticospinal excitability after 1800 pulses of subthreshold
motor cortex during viewing of speech-related lip 5 Hz repetitive TMS to the primary motor cortex. Clinical
movements or listening to speech. Neuropsychologia, Neurophysiology, 115, 1519–1526.
49, 2045–2054. Peterchev, A. V., Wagner, T. A., Miranda, P. C., Nitsche, M. A.,
Murakami, T., Restle, J., & Ziemann, U. (2012). Effective Paulus, W., Lisanby, S. H., et al. (2012). Fundamentals of
connectivity hierarchically links temporoparietal and frontal transcranial electric and magnetic stimulation dose:
areas of the auditory dorsal stream with the motor cortex lip Definition, selection, and reporting practices. Brain
area during speech perception. Brain and Language, 122, Stimulation, 5, 435–453.
135–141. Polania, R., Paulus, W., & Nitsche, M. A. (2012). Modulating
Nagel, T. (1993). What is the mind–body problem? Ciba cortico-striatal and thalamo-cortical functional connectivity
Foundation Symposium, 174, 1–7. with transcranial direct current stimulation. Human Brain
Neri, F., Mencarelli, L., Menardi, A., Giovannelli, F., Rossi, S., Mapping, 33, 2499–2508.
Sprugnoli, G., et al. (2020). A novel tDCS sham approach Price, C. J., & Friston, K. J. (2002). Degeneracy and cognitive
based on model-driven controlled shunting. Brain anatomy. Trends in Cognitive Sciences, 6, 416–421.
Stimulation, 13, 507–516. Rahman, A., Reato, D., Arlotti, M., Gasca, F., Datta, A., Parra, L. C.,
Neuling, T., Rach, S., & Herrmann, C. S. (2013). Orchestrating et al. (2013). Cellular effects of acute direct current
neuronal networks: Sustained after-effects of transcranial stimulation: Somatic and synaptic terminal effects. Journal of
alternating current stimulation depend upon brain states. Physiology, 591, 2563–2578.
Frontiers in Human Neuroscience, 7, 161. Ramsey, J. D., Hanson, S. J., Hanson, C., Halchenko, Y. O.,
Nguyen, J., Deng, Y., & Reinhart, R. M. G. (2018). Brain-state Poldrack, R. A., & Glymour, C. (2010). Six problems for causal
determines learning improvements after transcranial inference from fMRI. Neuroimage, 49, 1545–1558.
alternating-current stimulation to frontal cortex. Brain Ratcliff, R. (1978). A theory of memory retrieval. Psychological
Stimulation, 11, 723–736. Review, 85, 59–108.
Nitsche, M. A., Cohen, L. G., Wassermann, E. M., Priori, A., Lang, Ratcliff, R., & McKoon, G. (2008). The diffusion decision model:
N., Antal, A., et al. (2008). Transcranial direct current Theory and data for two-choice decision tasks. Neural
stimulation: State of the art 2008. Brain Stimulation, Computation, 20, 873–922.
1, 206–223. Ratcliff, R., Smith, P. L., Brown, S. D., & McKoon, G. (2016).
Nitsche, M. A., & Paulus, W. (2000). Excitability changes Diffusion decision model: Current issues and history. Trends
induced in the human motor cortex by weak transcranial in Cognitive Sciences, 20, 260–281.
direct current stimulation. Journal of Physiology, 527, Reato, D., Rahman, A., Bikson, M., & Parra, L. C. (2010). Low-
633–639. intensity electrical stimulation affects network dynamics by
Noury, N., Hipp, J. F., & Siegel, M. (2016). Physiological modulating population rate and spike timing. Journal of
processes non-linearly affect electrophysiological recordings Neuroscience, 30, 15067–15079.
during transcranial electric stimulation. Neuroimage, 140, Reato, D., Rahman, A., Bikson, M., & Parra, L. C. (2013). Effects
99–109. of weak transcranial alternating current stimulation on brain

26 Journal of Cognitive Neuroscience Volume X, Number Y


activity—A review of known mechanisms from animal uncertainty analysis in electric field calculations.
studies. Frontiers in Human Neuroscience, 7, 687. Neuroimage, 188, 821–834.
Reid, A. T., Headley, D. B., Mill, R. D., Sanchez-Romero, R., Scheines, R. (2005). The similarity of causal inference in
Uddin, L. Q., Marinazzo, D., et al. (2019). Advancing experimental and non-experimental studies. Philosophy of
functional connectivity research from association to Science, 72, 927–940.
causation. Nature Neuroscience, 22, 1751–1760. Schmid, U. D., Møller, A. R., & Schmid, J. (1995). Transcranial
Reinhart, R. M. G., & Nguyen, J. A. (2019). Working memory magnetic stimulation of the trigeminal nerve: Intraoperative
revived in older adults by synchronizing rhythmic brain study on stimulation characteristics in man. Muscle & Nerve,
circuits. Nature Neuroscience, 22, 820–827. 18, 487–494.
Ridderinkhof, K. R. (2002a). Activation and suppression in Schneidman, E., Puchalla, J. L., Segev, R., Harris, R. A., Bialek,
conflict tasks: Empirical clarification through distributional W., & Berry, M. J., II (2011). Synergy from silence in a
analyses. In W. Prinz & B. Hommel (Eds.), Attention and combinatorial neural code. Journal of Neuroscience, 31,
performance XIX: Common mechanisms in perception and 15732–15741.
action (pp. 494–519). Oxford, UK: Oxford University Press. Schroeder, C. E., & Lakatos, P. (2009). Low-frequency neuronal
Ridderinkhof, K. R. (2002b). Micro- and macro-adjustments of oscillations as instruments of sensory selection. Trends in
task set: Activation and suppression in conflict tasks. Neurosciences, 32, 9–18.
Psychological Research, 66, 312–323. Schuhmann, T., Schiller, N. O., Goebel, R., & Sack, A. T. (2012).
Ridding, M. C., & Ziemann, U. (2010). Determinants of the Speaking of which: Dissecting the neurocognitive network of
induction of cortical plasticity by non-invasive brain language production in picture naming. Cerebral Cortex, 22,
stimulation in healthy subjects. Journal of Physiology, 701–709.
588, 2291–2304. Schutter, D. J. L. G. (2016). Cutaneous retinal activation and
Riddle, J., Hwang, K., Cellier, D., Dhanani, S., & D’Esposito, M. neural entrainment in transcranial alternating current
(2019). Causal evidence for the role of neuronal oscillations stimulation: A systematic review. Neuroimage, 140, 83–88.
in top–down and bottom–up attention. Journal of Cognitive Schwarzkopf, D. S., Silvanto, J., & Rees, G. (2011). Stochastic
Neuroscience, 31, 768–779. resonance effects reveal the neural mechanisms of
Rogasch, N. C., Sullivan, C., Thomson, R. H., Rose, N. S., Bailey, transcranial magnetic stimulation. Journal of Neuroscience,
N. W., Fitzgerald, P. B., et al. (2017). Analysing concurrent 31, 3143–3147.
transcranial magnetic stimulation and Seghier, M. L. (2013). The angular gyrus: Multiple functions and
electroencephalographic data: A review and introduction to multiple subdivisions. Neuroscientist, 19, 43–61.
the open-source TESA software. Neuroimage, 147, 934–951. Shannon, C. E., & Weaver, W. (1949). The mathematical theory
Romei, V., Bauer, M., Brooks, J. L., Economides, M., Penny, W., of communication. Urbana, IL: University of Illinois Press.
Thut, G., et al. (2016). Causal evidence that intrinsic beta- Siebner, H. R. (2020). Does TMS of the precentral motor hand
frequency is relevant for enhanced signal propagation in the knob primarily stimulate the dorsal premotor cortex or the
motor system as shown through rhythmic TMS. Neuroimage, primary motor hand area? Brain Stimulation, 13, 517–518.
126, 120–130. Siebner, H. R., Auer, C., Roeck, R., & Conrad, B. (1999).
Romero, M. C., Davare, M., Armendariz, M., & Janssen, P. Trigeminal sensory input elicited by electric or magnetic
(2019). Neural effects of transcranial magnetic stimulation at stimulation interferes with the central motor drive to the
the single-cell level. Nature Communications, 10, 2642. intrinsic hand muscles. Clinical Neurophysiology, 110,
Rossi, S., Ferro, M., Cincotta, M., Ulivelli, M., Bartalini, S., 1090–1099.
Miniussi, C., et al. (2007). A real electro-magnetic placebo Siebner, H. R., Bergmann, T. O., Bestmann, S., Massimini, M.,
(REMP) device for sham transcranial magnetic stimulation Johansen-Berg, H., Mochizuki, H., et al. (2009). Consensus
(TMS). Clinical Neurophysiology, 118, 709–716. paper: Combining transcranial stimulation with
Ruff, C. C., Bestmann, S., Blankenburg, F., Bjoertomt, O., neuroimaging. Brain Stimulation, 2, 58–80.
Josephs, O., Weiskopf, N., et al. (2008). Distinct causal Siebner, H. R., Hartwigsen, G., Kassuba, T., & Rothwell, J. C.
influences of parietal versus frontal areas on human visual (2009). How does transcranial magnetic stimulation modify
cortex: Evidence from concurrent TMS–fMRI. Cerebral neuronal activity in the brain? Implications for studies of
Cortex, 18, 817–827. cognition. Cortex, 45, 1035–1042.
Ruzzoli, M., Marzi, C. A., & Miniussi, C. (2010). The neural Silvanto, J., Bona, S., Marelli, M., & Cattaneo, Z. (2018). On the
mechanisms of the effects of transcranial magnetic mechanisms of transcranial magnetic stimulation (TMS):
stimulation on perception. Journal of Neurophysiology, 103, How brain state and baseline performance level determine
2982–2989. behavioral effects of TMS. Frontiers in Psychology, 9, 741.
Sack, A. T. (2006). Transcranial magnetic stimulation, causal Silvanto, J., & Cattaneo, Z. (2017). Common framework for
structure–function mapping and networks of functional “virtual lesion” and state-dependent TMS: The
relevance. Current Opinion in Neurobiology, 16, 593–599. facilitatory/suppressive range model of online TMS effects on
Sack, A. T., Cohen Kadosh, R., Schuhmann, T., Moerel, M., behavior. Brain and Cognition, 119, 32–38.
Walsh, V., & Goebel, R. (2009). Optimizing functional Silvanto, J., Lavie, N., & Walsh, V. (2005). Double dissociation of
accuracy of TMS in cognitive studies: A comparison of V1 and V5/MT activity in visual awareness. Cerebral Cortex,
methods. Journal of Cognitive Neuroscience, 21, 207–221. 15, 1736–1741.
Sandrini, M., Umiltà, C., & Rusconi, E. (2011). The use of Silvanto, J., Muggleton, N., & Walsh, V. (2008). State-
transcranial magnetic stimulation in cognitive neuroscience: dependency in brain stimulation studies of perception and
A new synthesis of methodological issues. Neuroscience & cognition. Trends in Cognitive Sciences, 12, 447–454.
Biobehavioral Reviews, 35, 516–536. Silvanto, J., Muggleton, N. G., Cowey, A., & Walsh, V. (2007).
Saturnino, G. B., Antunes, A., & Thielscher, A. (2015). On the Neural adaptation reveals state-dependent effects of
importance of electrode parameters for shaping electric field transcranial magnetic stimulation. European Journal of
patterns generated by tDCS. Neuroimage, 120, 25–35. Neuroscience, 25, 1874–1881.
Saturnino, G. B., Thielscher, A., Madsen, K. H., Knösche, T. R., Smalle, E. H. M., Rogers, J., & Möttönen, R. (2015). Dissociating
& Weise, K. (2019). A principled approach to conductivity contributions of the motor cortex to speech perception and

Bergmann and Hartwigsen 27


response bias by using transcranial magnetic stimulation. activity and associated functions: A position paper. Clinical
Cerebral Cortex, 25, 3690–3698. Neurophysiology, 128, 843–857.
Snowball, A., Tachtsidis, I., Popescu, T., Thompson, J., Delazer, Thut, G., Schyns, P. G., & Gross, J. (2011). Entrainment of
M., Zamarian, L., et al. (2013). Long-term enhancement of perceptually relevant brain oscillations by non-invasive
brain function and cognition using cognitive training and rhythmic stimulation of the human brain. Frontiers in
brain stimulation. Current Biology, 23, 987–992. Psychology, 2, 170.
Sparing, R., Buelte, D., Meister, I. G., Pauš, T., & Fink, G. R. Thut, G., Veniero, D., Romei, V., Miniussi, C., Schyns, P., &
(2008). Transcranial magnetic stimulation and the challenge Gross, J. (2011). Rhythmic TMS causes local entrainment of
of coil placement: A comparison of conventional and natural oscillatory signatures. Current Biology, 21,
stereotaxic neuronavigational strategies. Human Brain 1176–1185.
Mapping, 29, 82–96. Tononi, G. (2008). Consciousness as integrated information: A
Sparing, R., Mottaghy, F. M., Ganis, G., Thompson, W. L., provisional manifesto. Biological Bulletin, 215, 216–242.
Töpper, R., Kosslyn, S. M., et al. (2002). Visual cortex Töpper, R., Mottaghy, F. M., Brügmann, M., Noth, J., & Huber,
excitability increases during visual mental imagery—A TMS W. (1998). Facilitation of picture naming by focal transcranial
study in healthy human subjects. Brain Research, 938, magnetic stimulation of Wernicke’s area. Experimental
92–97. Brain Research, 121, 371–378.
Spieser, L., van den Wildenberg, W., Hasbroucq, T., Turi, Z., Csifcsák, G., Boayue, N. M., Aslaksen, P., Antal, A.,
Ridderinkhof, K. R., & Burle, B. (2015). Controlling your Paulus, W., et al. (2019). Blinding is compromised for
impulses: Electrical stimulation of the human supplementary transcranial direct current stimulation at 1 mA for 20 min in
motor complex prevents impulsive errors. Journal of young healthy adults. European Journal of Neuroscience,
Neuroscience, 35, 3010–3015. 50, 3261–3268.
Stagg, C. J., & Nitsche, M. A. (2011). Physiological basis of Turi, Z., Paulus, W., & Antal, A. (2012). Functional neuroimaging
transcranial direct current stimulation. Neuroscientist, 17, and transcranial electrical stimulation. Clinical EEG and
37–53. Neuroscience, 43, 200–208.
Stefan, K., Kunesch, E., Cohen, L. G., Benecke, R., & Classen, J. Valls-Solé, J., Pascual-Leone, A., Wassermann, E. M., & Hallett,
(2000). Induction of plasticity in the human motor cortex by M. (1992). Human motor evoked responses to paired
paired associative stimulation. Brain, 123, 572–584. transcranial magnetic stimuli. Electroencephalography and
Stefan, K., Wycislo, M., Gentner, R., Schramm, A., Naumann, M., Clinical Neurophysiology, 85, 355–364.
Reiners, K., et al. (2006). Temporary occlusion of associative van Campen, A. D., Kunert, R., van den Wildenberg, W. P. M., &
motor cortical plasticity by prior dynamic motor training. Ridderinkhof, K. R. (2018). Repetitive transcranial magnetic
Cerebral Cortex, 16, 376–385. stimulation over inferior frontal cortex impairs the
Stetkarova, I., & Kofler, M. (2013). Differential effect of baclofen suppression (but not expression) of action impulses during
on cortical and spinal inhibitory circuits. Clinical action conflict. Psychophysiology, 55, e13003.
Neurophysiology, 124, 339–345. van den Wildenberg, W. P. M., Wylie, S. A., Forstmann, B. U.,
Stokes, M. G., Chambers, C. D., Gould, I. C., Henderson, T. R., Burle, B., Hasbroucq, T., & Ridderinkhof, K. R. (2010). To
Janko, N. E., Allen, N. B., et al. (2005). Simple metric for head or to heed? Beyond the surface of selective action
scaling motor threshold based on scalp–cortex distance: inhibition: A review. Frontiers in Human Neuroscience,
Application to studies using transcranial magnetic 4, 222.
stimulation. Journal of Neurophysiology, 94, 4520–4527. Vollmann, H., Conde, V., Sewerin, S., Taubert, M., Sehm, B.,
Taylor, P. C. J., Nobre, A. C., & Rushworth, M. F. S. (2007). FEF Witte, O. W., et al. (2013). Anodal transcranial direct current
TMS affects visual cortical activity. Cerebral Cortex, 17, stimulation (tDCS) over supplementary motor area (SMA)
391–399. but not pre-SMA promotes short-term visuomotor learning.
Terao, Y., Ugawa, Y., Suzuki, M., Sakai, K., Hanajima, R., Gemba- Brain Stimulation, 6, 101–107.
Shimizu, K., et al. (1997). Shortening of simple reaction time Volman, I., Roelofs, K., Koch, S., Verhagen, L., & Toni, I. (2011).
by peripheral electrical and submotor-threshold magnetic Anterior prefrontal cortex inhibition impairs control over
cortical stimulation. Experimental Brain Research, 115, social emotional actions. Current Biology, 21, 1766–1770.
541–545. Vöröslakos, M., Takeuchi, Y., Brinyiczki, K., Zombori, T., Oliva,
ter Braack, E. M., de Vos, C. C., & van Putten, M. J. A. M. (2015). A., Fernández-Ruiz, A., et al. (2018). Direct effects of
Masking the auditory evoked potential in TMS–EEG: A transcranial electric stimulation on brain circuits in rats and
comparison of various methods. Brain Topography, 28, humans. Nature Communications, 9, 483.
520–528. Voss, A., Nagler, M., & Lerche, V. (2013). Diffusion models in
Terney, D., Chaieb, L., Moliadze, V., Antal, A., & Paulus, W. experimental psychology: A practical introduction.
(2008). Increasing human brain excitability by transcranial Experimental Psychology, 60, 385–402.
high-frequency random noise stimulation. Journal of Voss, A., & Voss, J. (2007). Fast-dm: A free program for efficient
Neuroscience, 28, 14147–14155. diffusion model analysis. Behavior Research Methods, 39,
Thielscher, A., & Kammer, T. (2004). Electric field properties of 767–775.
two commercial figure-8 coils in TMS: Calculation of focality Vossen, A., Gross, J., & Thut, G. (2015). Alpha power increase
and efficiency. Clinical Neurophysiology, 115, 1697–1708. after transcranial alternating current stimulation at alpha
Thielscher, A., Opitz, A., & Windhoff, M. (2011). Impact of the frequency (alpha-tACS) reflects plastic changes rather than
gyral geometry on the electric field induced by transcranial entrainment. Brain Stimulation, 8, 499–508.
magnetic stimulation. Neuroimage, 54, 234–243. Vosskuhl, J., Strüber, D., & Herrmann, C. S. (2018). Non-
Thies, M., Zrenner, C., Ziemann, U., & Bergmann, T. O. (2018). invasive brain stimulation: A paradigm shift in understanding
Sensorimotor mu-alpha power is positively related to brain oscillations. Frontiers in Human Neuroscience,
corticospinal excitability. Brain Stimulation, 11, 1119–1122. 12, 211.
Thut, G., Bergmann, T. O., Fröhlich, F., Soekadar, S. R., Brittain, Walsh, V., & Cowey, A. (2000). Transcranial magnetic
J.-S., Valero-Cabré, A., et al. (2017). Guiding transcranial brain stimulation and cognitive neuroscience. Nature Reviews
stimulation by EEG/MEG to interact with ongoing brain Neuroscience, 1, 73–80.

28 Journal of Cognitive Neuroscience Volume X, Number Y


Walsh, V., Ellison, A., Battelli, L., & Cowey, A. (1998). Task- Womelsdorf, T., Valiante, T. A., Sahin, N. T., Miller, K. J., &
specific impairments and enhancements induced by Tiesinga, P. (2014). Dynamic circuit motifs underlying
magnetic stimulation of human visual area V5. Proceedings of rhythmic gain control, gating and integration. Nature
the Royal Society of London, Series B, Biological Sciences, Neuroscience, 17, 1031–1039.
265, 537–543. Yavari, F., Jamil, A., Mosayebi Samani, M., Vidor, L. P., & Nitsche,
Ward, N. S., Bestmann, S., Hartwigsen, G., Weiss, M. M., M. A. (2018). Basic and functional effects of transcranial
Christensen, L. O. D., Frackowiak, R. S. J., et al. (2010). Low- electrical stimulation (tES)—An introduction. Neuroscience
frequency transcranial magnetic stimulation over left dorsal & Biobehavioral Reviews, 85, 81–92.
premotor cortex improves the dynamic control of visuospatially Zaehle, T., Rach, S., & Herrmann, C. S. (2010). Transcranial
cued actions. Journal of Neuroscience, 30, 9216–9223. alternating current stimulation enhances individual alpha
Wassermann, E. M., Blaxton, T. A., Hoffman, E. A., Berry, C. D., activity in human EEG. PLoS One, 5, e13766.
Oletsky, H., Pascual-Leone, A., et al. (1999). Repetitive Zazio, A., Bortoletto, M., Ruzzoli, M., Miniussi, C., & Veniero, D.
transcranial magnetic stimulation of the dominant (2019). Perceptual and physiological consequences of dark
hemisphere can disrupt visual naming in temporal lobe adaptation: A TMS–EEG study. Brain Topography, 32,
epilepsy patients. Neuropsychologia, 37, 537–544. 773–782.
Weichwald, S., Meyer, T., Özdenizci, O., Schölkopf, B., Ball, T., Ziemann, U., Paulus, W., Nitsche, M. A., Pascual-Leone, A.,
& Grosse-Wentrup, M. (2015). Causal interpretation rules for Byblow, W. D., Berardelli, A., et al. (2008). Consensus: Motor
encoding and decoding models in neuroimaging. cortex plasticity protocols. Brain Stimulation, 1, 164–182.
Neuroimage, 110, 48–59. Zimmermann, M., Verhagen, L., de Lange, F. P., & Toni, I.
Weise, K., Numssen, O., Thielscher, A., Hartwigsen, G., & (2016). The extrastriate body area computes desired goal
Knösche, T. R. (2020). A novel approach to localize cortical states during action planning. eNeuro, 3, ENEURO.0020-
TMS effects. Neuroimage, 209, 116486. 16.2016.

Bergmann and Hartwigsen 29

You might also like